You are on page 1of 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/327065847

Comparative Study of State-of-the-Art Macroscopic Models for Planar


Reinforced Concrete Walls

Article  in  Aci Structural Journal · November 2018


DOI: 10.14359/51710835

CITATIONS READS

11 1,434

10 authors, including:

Kristijan Kolozvari Carlos A Arteta


California State University, Fullerton Universidad del Norte (Colombia)
16 PUBLICATIONS   167 CITATIONS    29 PUBLICATIONS   46 CITATIONS   

SEE PROFILE SEE PROFILE

Matias Hube
Pontificia Universidad Católica de Chile
71 PUBLICATIONS   327 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CEER - Thin Wall Project View project

Forecast Engineering: From Past Design to Future Decisions | ERASMUS+ Programme - Key Activity 2: Strategic Partnerships in Higher Education View project

All content following this page was uploaded by John W. Wallace on 17 June 2019.

The user has requested enhancement of the downloaded file.


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title No. 115-S127

Comparative Study of State-of-the-Art Macroscopic Models


for Planar Reinforced Concrete Walls
by K. Kolozvari, C. Arteta, M. Fischinger, S. Gavridou, M. Hube, T. Isaković, L. Lowes,
K. Orakcal, J. Vásquez, and J. Wallace
Over the past 20 years, a spectrum of analytical models for beam models with concentrated plastic hinges generally
nonlinear analysis of reinforced concrete (RC) structural walls, with provides reasonably accurate results for beams, while the
varying capabilities and complexities, have become available for same approach is generally less accurate for simulation
both research and design applications. Five conceptually different of column responses (due to axial/flexural interaction) or
state-of-the-art macroscopic models were described, including
structural wall responses (due to axial/flexural-shear inter-
two-node and four-node elements, based on either a fiber-based
action and neutral axis migration). Therefore, it is essential
representation of a wall cross section or a strut-and-tie approach,
using either force-deformation or strain-stress material behavior, to understand the capabilities and limitations of numerical
and considering either coupled or uncoupled axial/flexural and models used to assess nonlinear structural responses so that
shear responses. Modeling approaches were validated against the results obtained can be appropriately interpreted.
experimental data obtained for five RC wall specimens character- Use of reinforced concrete (RC) structural walls is very
ized by a range of properties (for example, aspect ratio, axial load, common for design of new buildings or retrofit of existing
and failure mechanism) to assess current modeling capabilities buildings, and the ability to accurately model their seismic
and identify future research directions. Results presented suggest behavior is essential. Despite this need, a vast majority of
that the considered analytical models typically overestimate initial model validation studies available in the literature (Orakcal
wall stiffness, models with uncoupled flexural and shear behavior and Wallace 2006; Massone et al. 2006; Jiang and Kurama
overestimate lateral capacity of walls where shear deformations
2010; Fischinger et al. 2012; Kolozvari et al. 2015a,b)
are significant, models with shear-flexure interaction can capture
typically report results from a single test program with
nonlinear shear deformations, and that vertical strains within the
plastic hinge region can be either overestimated or underestimated limited test variables and an analytical modeling approach
by a factor of 2 in the nonlinear response range using the plane implemented that is tailored to assess the impacts of the
sections assumption. given test program; studies that focus on a more compre-
hensive assessment of available models and the ability of
Keywords: macroscopic models; nonlinear modeling; performance-based these models to capture salient responses from a broad set
design; reinforced concrete walls. of test programs and conditions are rare (Pugh et al. 2015;
Baker et al. 2016). Prior to 1990, relatively few tests of RC
INTRODUCTION walls were reported in the literature, and a majority of tests
Application of nonlinear structural analysis procedures to focused on behavior of walls with well-confined boundary
assess the expected seismic performance of existing build- columns at the wall edges and a relatively thin wall web
ings (for example, using ASCE 41-17 [2017]) or for design (Oesterle et al. 1976, 1979; Vallenas et al. 1979), which was
of new buildings (for example, using Los Angeles Tall in line with general U.S. practice at the time. After 1990,
Buildings Structural Design Council (LATBSDC) [2017]) a large number of wall tests were designed and conducted
has become common in regions of moderate-to-high seis- to address a wide range of issues, including flexure-domi-
micity. Use of nonlinear building models subjected to ground nant behavior (for example, slender walls with low shear;
acceleration histories generally allows for a more reliable M/Vlw > 3.0), shear-dominant behavior (for example, squat
assessment of system and element demands (for example, walls with high shear; M/Vlw < 1.0), and behavior of walls
lateral interstory drift, wall shear stress, or local strains or where both flexure and shear responses may be important
rotations), which are then compared with defined limits to (for example, 1.0 < M/Vlw < 3.0). Most of the post-1990s test
assess if acceptable performance is expected. The accuracy results are well-documented in written reports, while detailed
and robustness of analytical approaches used to simulate digital data and photos/figures of wall behavior and failure
the nonlinear behavior of structural components will affect mechanisms are typically available in web-based experi-
analysis results, and ultimately the structural design; there- mental databases (for example, NHERI 2018; NEES 2016).
fore, it is essential that reliable computational models are These data generally include information on global and local
available. With expansion of computational capabilities over
the past 20 years, a vast number of analytical approaches
with a broad range of capabilities have been developed based
ACI Structural Journal, V. 115, No. 6, November 2018.
on various theories, and are used in both research and engi- MS No. S-2017-269, doi: 10.14359/51710835, was received July 24, 2017, and
neering practice. The accuracy and range of applicability reviewed under Institute publication policies. Copyright © 2018, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
of available modeling approaches vary greatly depending obtained from the copyright proprietors. Pertinent discussion including author’s
closure, if any, will be published ten months from this journal’s date if the discussion
on how the models are applied. For example, use of elastic is received within four months of the paper’s print publication.

ACI Structural Journal/November 2018 1637


responses that are essential for comprehensive assessments on a layered membrane finite element formulation capable
of modeling approaches available in the literature. of capturing lateral instabilities at wall boundaries, a failure
Analytical modeling of the inelastic response of struc- mode observed following recent earthquakes in Chile in
tural walls can be accomplished by either using microscopic 2010 and New Zealand in 2011 and in recent tests (Wallace
or macroscopic models. Microscopic (for example, finite 2012), have recently been developed and validated against
element) models can provide a refined and detailed estima- limited test data (Dashti et al. 2014). Finally, availability of
tion of structural responses, but significant effort is  typi- a common, research-oriented, and modular computational
cally required to develop the model, perform the analysis platform, namely OpenSees (McKenna et al. 2000), as well
(computational effort), and interpret the results. For these as other computational platforms (for example, LS-Dyna,
reasons, use of micro-models is somewhat impractical for VecTor), have led to significant advances in numerical simu-
engineering design. In contrast, macro-element models lation capabilities to model walls and wall systems for both
(or macroscopic models) are effective computational tools researchers and practicing engineers.
as they are relatively simple to implement and have been As summarized in the prior two paragraphs, over the past
shown to be reasonably accurate and efficient in predicting 20 years, and especially in the last 10 years, structural-earth-
the hysteretic response of RC structural walls (Taucer et quake engineering has seen major advances in the avail-
al. 1991; Computers and Structures Inc. 2005; Orakcal and ability of robust wall test data sets as well as in the diversity
Wallace 2006; Jiang and Kurama 2010; Fischinger et al. and sophistication of wall modeling tools. Contributions
2014; Kolozvari et al. 2015a; Vásquez et al. 2016). However, have come from all over the globe, with significant studies
to obtain reliable results using macro models, one should reported in the United States, Chile, New Zealand, Europe,
carefully choose model and material parameters within the and Japan, as well as other countries. The results reported
range of interest for a given application, assess sensitivity here were part of an organized international effort to: 1)
of the results to variations in material and modeling param- share test results; 2) advance modeling approaches; 3) iden-
eters, and understand model assumptions and limitations. tify research gaps; and 4) promote/improve collaboration.
A large number of macroscopic models for RC walls have
become available over the past two decades that are char- SCOPE AND OBJECTIVES
acterized with various assumptions used in model devel- Given the significant number of macroscopic analyt-
opment, and therefore, various ranges of applicability. The ical models for nonlinear analysis of RC walls available in
most commonly used macroscopic modeling approaches various computational platforms and reported in the litera-
simulate wall flexural behavior using force-displacement ture, as well as their wide application in both engineering
or strain-stress material models for axial-flexural behavior practice and research, there is a need for a comprehensive
and shear behavior with a single horizontal spring using a overview of such models and assessment of their capabil-
simplified hysteresis response (for example, linear-elastic, ities. State-of-the-art wall macro-modeling approaches
elasto-plastic, origin oriented). Typically, shear behavior is reported in the literature are systematically evaluated here
uncoupled (or independent) from the axial-bending (flexural) by comparing responses computed with the models with
behavior; therefore, it is common to refer to these models as responses obtained in various tests programs on isolated
“uncoupled” models. However, interaction between axial/ planar RC walls to provide valuable information on model
flexural and shear behavior in RC was observed in various capabilities and limitations, as well as to identify gaps that
experiments (Oesterle et al. 1976; Tran and Wallace 2015), require additional research and prioritize research needs.
where Massone et al. (2006) and Fischinger et al. (2017) This effort was made possible by engaging a broad group of
demonstrated that this interaction could be important even in international researchers to conduct coordinated modeling
slender RC walls. Over the last decade, studies have focused studies for a common set of wall tests covering a range of
on extending macro models to couple axial-bending and behaviors reported in the literature. The models and the test
shear behavior to incorporate nonlinear shear-flexure inter- data are archived to encourage future studies that advance
action (Massone et al. 2006; Jiang and Kurama 2010; Fisch- the work presented. Specific objectives of this paper are to:
inger et al. 2014; Kolozvari et al. 2015a); these models are 1. Provide a state-of-the-art review for a range of wall
commonly referred to as “coupled” models. Considerable macroscopic models used to assess the nonlinear behavior
effort has also been expended on improving macro models to of RC structural walls;
incorporate mechanisms that lead to loss of wall lateral load 2. Conduct a systematic study for considered macroscopic
capacity (for example, reinforcement buckling: Gomes and modeling approaches to enable comparisons between model
Appleton 1997; Dhakal and Maekawa 2002b) and axial load and test results for both global and local responses; and
capacity (Wallace et al. 2008) and to implement these models 3. Assess modeling approaches with respect to objectivity,
into some computational platforms (McKenna et al. 2000). ease of use, range of applicability, and accuracy within that
However, the ability of these models to capture strength range of applicability.
degradation observed in tests of isolated structural walls or Brief descriptions of the selected modeling approaches
system level tests that include structural walls has not been are presented first, followed by descriptions of the experi-
studied sufficiently. For example, models used to capture mental studies used to evaluate and compare model results.
reinforcement buckling (Dhakal and Maekawa 2002a) are Comparisons between experimentally measured and analyt-
based on tests of single reinforcing bars in air, versus groups ically predicted wall responses are then provided at both
of bars embedded in concrete. Modeling approaches based global and local response levels for all modeling approaches

1638 ACI Structural Journal/November 2018


considered. Based on the results presented, conclusions are
drawn regarding model capabilities and limitations.

DESCRIPTION OF MACROSCOPIC MODELING


APPROACHES
Five conceptually different macroscopic modeling
approaches for RC walls are selected for this study:
1. Shear wall element implemented in the commercial
software Perform 3D (P3D-SW [Computers and Structures
Inc. 2005]);
2. Force wall element (FWE [Vásquez et al. 2016])
derived from the force-based nonlinear fiber-section beam-
column element;
3. Multiple Vertical Line Element Model (MVLEM) with
material behavior defined by simplified force-deformation
relationships without shear-flexure interaction (MVLEM-FD
[Fischinger et al. 2004]) and with shear-flexural interaction
(SFI-MVLEM-FD [Fischinger et al. 2012]) implemented in
the University of Ljubljana’s in-house version of OpenSees;
4. Shear-Flexure Interaction Multiple Vertical Line
Element Model available in OpenSees, with material Fig. 1—Shear wall element: (a) vertical axial/bending; (b)
behavior defined using uniaxial strain-stress relationships concrete shear; and (c) typical material backbone relation
(SFI-MVLEM-SS [Kolozvari et al. 2015a, 2018]); and (Computers and Structures Inc. 2005).
5. Nonlinear Truss Model (NLTM [Panagiotou et al.
2012]) also available in OpenSees, based on a strut-and-tie (Fig. 1(b)). In-plane shear and flexural responses are uncou-
modeling approach. pled, while out-of-plane shear and bending behavior are
The analytical models selected in this study represent defined as linear-elastic.
state-of-the-art macro modeling approaches used in research Material models for concrete, steel, and shear behavior
and practice to assess nonlinear responses of RC structural are defined using the generic backbone relation presented
walls. The following sections provide brief overview of the in Fig. 1(c), which is characterized by strain and stress (or
theory underlying the analytical models considered, with the deformation and force) values corresponding to yield (Y),
emphasis on their advantages and limitations. ultimate (U), strength loss (L), residual (R), and maximum
deformation (X). In addition, the constitutive model uses
PERFORM-3D shear wall element simplified unloading/reloading rules that cannot capture
PERFORM-3D (Computers and Structures Inc. 2005) is detailed material hysteretic response (for example, gradual
the most commonly used commercial software in the United gap closure in concrete, or the Bauschinger effect in rein-
States for nonlinear analysis of RC buildings. Backbone forcing steel). Nevertheless, the model is computationally
curves used to define the nonlinear behavior of structural stable and requires relatively low computational effort,
components available in PERFORM-3D are compatible with which makes it attractive for practical applications.
the backbone relationships published in ASCE 41-17 (2017),
which allows direct implementation of performance-based Distributed plasticity fiber-based beam-column
design methodologies in engineering practice. Perform 3D force wall element
offers two types of elements for nonlinear analysis of RC Distributed plasticity fiber-based beam-column model
walls: 1) shear-wall element; and 2) general wall element. elements are available in the literature in the form of
Although formulation of the shear wall element represents a displacement-based and force-based element formulations
simplification of the general wall element, it has become the (Spacone et al. 1996; Neuenhofer and Filippou 1998; Taucer
more commonly used element for simulation of RC walls in et al. 1991; Scott et al. 2004). As an example of the distrib-
Perform 3D. A brief description of the shear wall element is uted plasticity beam-column model element, the force-based
provided in the following paragraphs, whereas details can force wall element (FWE) proposed by Vásquez et al. (2016)
be found in the Perform 3D user manual (Computers and is investigated in this study.
Structures Inc. 2005). Similarly to other formulations of fiber-based models,
The shear wall element in Perform 3D (P3D-SW) is in the FEW, axial and flexural behavior are simulated by a
defined as a four-node macroscopic model element, with series of longitudinal uniaxial fibers representing the wall
horizontal and vertical displacement degrees of freedom at cross section and the vertical reinforcement, along with the
each node. The behavior of the model element is governed by assumption that plane sections remain plane upon loading
two layers connected in parallel (Fig. 1): 1) a horizontal fiber (Fig. 2(a)). Stiffness and force-deformation properties of the
section used to represent axial/bending behavior of concrete longitudinal fibers are defined based on uniaxial stress-strain
and reinforcing steel in the vertical direction (Fig. 1(a)); relations for concrete and reinforcing steel, evaluated at a
and 2) a layer used to represent concrete shear behavior number of integration points (use of three to seven integra-

ACI Structural Journal/November 2018 1639


Fig. 2—Force-based fiber-section nonlinear beam-column element.

Fig. 3—MVLEM formulations: (a) 2-D version (Orakcal et al. 2004); and (b) 3-D version (Fischinger et al. 2004).
tion points per story is typical), and corresponding areas of is a major disadvantage of the FWE and other fiber-based
the materials on the wall cross section. Nodal resultants are models available in the literature.
computed using an appropriate numerical integration rule
along the length of the element. FWE employs material regu- Multiple-Vertical-Line-Element-Models with and
larization to obtain cyclic predictions of the wall behavior without shear-flexural interaction
that are insensitive to mesh size, particularly in the softening The Multiple-Vertical-Line-Element-Model (MVLEM)
region (that is, loss of capacity) of the wall load-deformation was originally proposed by Kabeyasawa et al. (1983) and
response (Coleman and Spacone 2001); details on material later modified and extended by many researchers (Vulcano et
regularization used in FWE are presented by Vásquez et al. al. 1989; Fischinger et al. 1990; Orakcal and Wallace 2006;
(2016). Chowdhury and Orakcal 2013). In the original MVLEM
Shear flexibility is incorporated in the FWE via shear formulation, axial/flexural behavior of a wall segment is
sections (Fig. 2(b)), which is common with other force-based simulated using a number of uniaxial springs (macro-fibers)
beam-column elements. Specific to the FWE, the behavior connected to rigid beams at the top and the bottom of the
of the shear section is defined using backbone relationship wall segments, as illustrated in Fig. 3 for two- and three-di-
by Gérin and Adebar (2009) that is characterized with four mensional (2-D and 3-D, resepectively) versions of the
points: 1) cracking (γcr, τcr), where shear cracking capacity is model. The shear response of the model element is uncou-
calculated using ACI 318-14 Equation 22.5.6.1; 2) yielding pled from the axial and flexural responses, and is defined
(γy, τy), which is not reached under flexural compression using a horizontal spring (Fig. 3(a) and (b)); the location
dominated behavior; 3) failure (γf, τf), which is only reached of the spring also represents the assumed center of rotation
in the section with largest demand, due to localization of for the element. In the presented study, two conceptually
deformations; and 4) the maximum shear resistance under different MVLEM-FD and MVLEM-SS were employed:
flexural compression behavior (γflex, τflex). In the FWE, shear 1) MVLEM-FD defines the response using force-displace-
response is dependent on axial load only by incorporating ments approach, and 2) MVLEM-SS is based on stress-
the Modified Compression Field Theory (Vecchio and strain relationships.
Collins 1986) at the shear section level. However, shear MVLEM-FD—In the basic version of MVLEM-FD, the
behavior is uncoupled from the flexural behavior, which force-displacement relationships presented in Fig. 4(a) and

1640 ACI Structural Journal/November 2018


Fig. 4—MVLEM-FD force-displacement responses of: (a) vertical springs; and (b) shear spring.

Fig. 5—SFI-MVLEM-FD accounting for inelastic shear and shear-flexural interaction in structural walls (vertical springs are
not shown).
(b) are used to define the flexural and shear response, respec- HSA mechanism in Specimens R2 and WSH6; for specimen
tively. MVLEM-FD was extended to account for axial-flex- descriptions, refer to the section “Experimental Date Used
ural-shear interaction (Rejec 2011) by introducing a hori- for Model Evaluation”. Factors that define pinching, damage
zontal shear spring on each vertical spring (macro-fiber), and degradation of the unloading stiffness in the Hysteretic
where the behavior of each horizontal and vertical spring model were specified based on typical shapes of hysteretic
is coupled (Fig. 5); this model is referred to as SFI-MV- loops (corresponding to HSA, HSD, and HSS mechanisms)
LEM-FD. Each of the horizontal springs takes into account as reported in the previous benchmark studies: α = 1.0, β =
three shear mechanisms: 1) dowel effect of vertical bars 1.5, γ = 1.05, δ = 0.5 (Fig. 4). More details about SFI-MV-
(HSD) described by relationships proposed by Dulacska LEM-FD can be found in Fischinger et al. (2014).
(1972) and Vintzeleou and Tassios (1987); 2) axial resis- MVLEM-SS—Another modification of the original
tance of horizontal/shear bars (HSS) defined using model by MVLEM to incorporate interaction between shear and flex-
Elwood and Moehle (2003); and 3) interlock of aggregate ural behavior was proposed by Kolozvari et al. (2015a,b;
particles in the crack (HSA) modeled using constitutive rela- Fig. 6) and implemented into publicly available version
tionships by Vecchio and Lai (2004). The current state of of the computational platform OpenSees (Kolozvari et al.
each spring component depends on deformations/displace- 2018). The proposed model, called the shear-flexure interac-
ments at the effective cracks of the element that are linked tion MVLEM (SFI-MVLEM-SS), incorporates biaxial RC
to the current displacements of the element nodes, which panel behavior into a two-dimensional formulation of the
enables coupling between axial/flexural and shear behavior MVLEM-SS implemented by Orakcal el al. (2004). Axial-
at the model element level. In the SFI-MVLEM-FD model shear coupling is achieved at the panel (macro-fiber) level,
formulation used in this study, constitutive behavior of the which allows coupling of axial/flexural and shear responses
shear resisting mechanisms (HSA, HSD, and HSS) are at the model element level. The constitutive RC panel model
represented with the Hysteretic material model available in implemented in the SFI-MVLEM-SS is the OpenSees mate-
OpenSees for Specimens RW-A15-P10-S78 and S6, while rial FSAM, which is an extension of the Fixed Strut Angle
the ShearSlip material (Kante 2005) is used to model the Model (Ulugtekin 2010, Orakcal et al. 2012; Fig. 6(b)).

ACI Structural Journal/November 2018 1641


Fig. 6—SFI-MVLE-SS formulation: (a) model element; (b) RC panel behavior; (c) concrete struts; (d) shear aggregate inter-
lock; (e) steel behavior; and (f) reinforcement dowel action.

Fig. 7—Hysteretic rules controlling response of springs in SFI-MVLEM-SS element: (a) strain-stress behavior of concrete; and
(b) strain-stress behavior of steel. (Note: 1 MPa = 0.145 ksi.)
The FSAM incorporates several strain-stress mechanisms concrete and steel are applied (Fig. 8(b)). The NLTM is
to represent the behavior of RC material including: 1) a implemented in the publicly available computational plat-
biaxial concrete behavior based on the fixed-strut model that form OpenSees as element Truss2 coupled with uniaxial
employs OpenSees uniaxial material ConcreteCM (Fig. 7(a)) material ConcretewBeta. In this study, the wall boundary
based on Chang and Mander (1994) formulation (Fig. 6(c)); elements (when present) are modeled using continuous
2) a uniaxial constitutive relationship for steel represented by nonlinear fiber-based beam-column elements (Spacone et al.
OpenSees material SteelMPF that follows the Menegotto and 1996), while remaining vertical and horizontal elements are
Pinto (1973) formulation (Fig. 7(b)) and is applied along the modeled as truss elements with fiber sections. The layout of
directions of horizontal and vertical reinforcement (Fig. 6(e)); diagonal truss elements resembles, but does not necessarily
3) a friction-based constitutive model used to capture shear match, the principal compressive stress trajectories when
aggregate interlock effects along concrete cracks (Fig. 6(d)); approaching the ultimate load, where inclination angle of the
and 4) a linear-elastic relationship representing dowel action diagonals proposed by Lu and Panagiotou (2014) is used.
on reinforcement (Fig. 6(f)). The shear spring is removed Strut angles used in this study for the wall specimens consid-
from the original MVLEM-SS formulation because the shear ered are as follows: 44.9 degress for RW2, 55.2 degrees for
stiffness and strength of each macro fiber (and therefore the SP4, 63.4 degrees for R2, 59.0 degrees for WSH6, and 48.7
model element) evolve according to the implemented consti- degrees for S6; for specimen descriptions, refer to the section
tutive RC panel behavior. Therefore, explicit definition of “Experimental Date Used for Model Evaluation”.
shear modeling parameters is not necessary. Each panel of concrete diagonals comprises two four-node
elements, where two of the nodes connect concrete diagonal
Nonlinear Truss Model elements to the main truss, while the others allow connecting
The Nonlinear Truss Model (NLTM) proposed by Panag- a virtual strain-gauge element to monitor the strains normal
iotou et al. (2012) is based on the strut-and-tie (truss) to the diagonal of interest (Fig. 9(a)). The capacity of the
approach, where a structural wall is modeled using nonlinear diagonal concrete truss elements is reduced as a function of
vertical, horizontal and diagonal truss elements connected transverse tensile strain, which allows coupling between the
at nodes along which cyclic uniaxial material laws for element compressive stress-strain behavior with the tensile

1642 ACI Structural Journal/November 2018


Fig. 8—NLTM modeling approach illustrated on Specimen SP4 (Tran and Wallace 2015). (Note: 1 in. = 25.4 mm.)
Summary of models
As described in the previous sections, all models consid-
ered in this study, except NLTM, are based on fiber repre-
sentation of reinforced concrete cross section of the wall,
implementation of the plane-sections-remain-plane assump-
tion, and two-node beam-column-type line element formu-
lation, with exception of P3D-SW, which is four-node
fiber-based element. In addition, most of the models can
be used to model walls with non-rectangular cross sections
and employed in three-dimensional response analysis. A
Fig. 9—(a) Four-node truss element formulation for concrete summary of major characteristics of modeling approaches
diagonal struts; and (b) relationship between concrete considered is presented in Table 1.
compressive stress reduction factor β and normal strain εn In modeling approaches that consider P-M interaction only,
(Lu et al. 2014). flexural behavior is simulated using uniaxial macro-fibers that
represent the behavior of concrete and reinforcing steel, the
strain normal to the element axis (Fig. 9(b)) and allows
behavior of which is defined using either uniaxial strain-stress
capturing the compression field of concrete (Vecchio and
or force-deformation hysteretic relationships. Shear behavior
Collins 1986). Therefore, NLTM captures interaction between
is described independently from flexural behavior via a shear
axial/flexural and shear behavior at the model element level,
spring (or shear section in the FWE), the behavior of which
and does not enforce plane section assumption, which distin-
is governed by a predefined hysteretic force-deformation rule.
guishes it from other macroscopic models presented here.
Forces and deformations are treated either at discrete integra-
However, due to the overlapping areas of vertical, horizontal
tion points or as average values over the element height, which
and diagonal concrete elements, precracking (initial) stiff-
is a trade-off between the ability of the model to describe local
ness and strength of walls are typically overestimated by
responses and numerical efficiency.
the model, as shown later in this study. As well, the shear
Two of the models that are capable of capturing axial/flexural
strength, shear stiffness, and the local element responses are
and shear interaction (SFI-MVLEM-FD and SFI-MVLEM-SS)
sensitive to the orientation of diagonal truss elements and
are extensions of previously developed uncoupled models, and
the element size suggesting that applicability of this model
incorporate interaction between axial and shear responses at the
may not be suitable for modeling of structural walls where
macro-fiber level via implementation of either a 2-D consti-
the principal compression stress orientations are expected to
tutive model for representing RC panel (membrane) behavior
vary significantly during seismic loading (for example, in
in each macro-fiber or by introducing a shear spring to each
coupled walls with varying axial loading). Finally, the large
macro-fiber that is coupled with axial fiber behavior. In these
number of degrees of freedom involved in generation of the
coupled modeling approaches, a critical component of the
NLTM wall model increases the computational demand.
model formulation is representation of the relatively complex

ACI Structural Journal/November 2018 1643


Table 1—Macroscopic models for reinforced concrete walls
Formulation Capabilities
Nodes per SS versus Average versus Plane sections Element 2-D versus Cross Publically
Model acronym element FD* discrete† remain plane? interactions 3-D analysis section availaible?
SW-P3D 4 SS Average Yes P-M 3-D Any PERFORM-3D
FWE 2 SS Discrete Yes P-M 2-D and 3-D Any No
MVLEM-FD ‡
2 FD Average Yes P-M 2-D and 3-D Any No§
SFI-MVLEM-FD 2 FD Average Yes P-M-V 2-D and 3-D Any No§
MVLEM-SS‡ 2 SS Average Yes P-M 2-D Planar OpenSees
SFI-MVLEM-SS 2 SS Average Yes P-M-V 2-D Planar OpenSees
NLTM 4 SS Average No P-M-V 3-D Any OpenSees
*
SS is strain-stress; FD is force-determination.

Forces and deformations at element level treated in average or discrete sense.

Limited or no results presented here.
§
Implemented into University of Ljubljana in-house version of OpenSees.

shear-resisting mechanisms along concrete cracks. Differently, section “Description of macroscopic modeling approaches,”
the NLTM incorporates axial/flexural and shear interaction via including discretization of model geometry, calibration of
a strut-and-tie modeling approach. material models, and load application.

EXPERIMENTAL DATA USED FOR MODEL Mesh information


EVALUATION The wall specimens were modeled with a suitable number
Experimental results obtained for five planar (rectangular) of macro-elements over the specimen height, considering
wall specimens subjected to in-plane lateral and axial loading the locations of the instrumentation used to measure defor-
were used to evaluate the modeling approaches considered mations on the wall specimens, to allow consistent local
in this study. The following five RC wall specimens were response comparisons between model predictions and exper-
investigated within the scope of this study: 1) RW2 (Thomsen imental results. Selected element heights varied between
and Wallace 1995); 2) R2 (Oesterle et al. 1976); 3) WSH6 one-fourth to one-half of the wall length to capture reason-
(Dazio et al. 2009); 4) RW-A15-P10-S78 (referred to as SP4 ably nonlinear deformations within the plastic hinge region,
for brevity; Tran and Wallace 2015); and 5) S6 (Vallenas et al. which is typically assumed to be approximately one-half of
1979). Specimens were selected to span a range of salient wall the wall length. Cross sections of the walls were divided into
response parameters and characteristics such as wall aspect a number of segments (macro-fibers) according to the place-
ratio (shear span ratio), axial load, shear demand at nominal ment and properties of the longitudinal reinforcing bars,
flexural capacity, and failure mode. The aspect ratio of the where wall boundary and web fibers were differentiated in
walls ranged from 1.50 to 3.13, where shear demand-capacity all modeling approaches. As an example, typical geometric
ratio was less than 1.0, including three walls with predom- discretization used in the modeling approaches considered
inantly flexural behavior and two specimens with shear- is illustrated in Fig. 10 for Specimen SP4, showing that
flexural interaction response observed during testing. Approx- for P3D-SW, SFI-MVLEM-FD, and SFI-MVLEM-SS, the
imate levels of axial load applied on the walls were 0% (one element heights of 305 and 406 mm (12 and 16 in.) were
specimen), 5% (one specimen), and 10% (three specimens) used; for NLTM, element heights of 366 mm (14.4 in.)
of wall axial load capacity. Observed failure modes included were used; and for FEW, element heights of 183, 996, and
flexure-dominated tension or compression failures (for 1300 mm (7.2, 39.2, and 51.2 in.) were adopted. Note that
example, bar fracture or concrete crushing and bar buckling) for FWE, a zero-length element is used at the base of the
of four specimens, and diagonal compression failure of one walls to account for strain penetration effects. A similar
specimen. For all specimens considered, data required to fully discretization methodology is used for all other specimens.
define a numerical model (that is, geometry, material prop- As illustrated in Fig. 10(a), discretization of the wall spec-
erties, reinforcement layout, and loading/boundary condi- imens for the P3D-SW model was performed using two
tions) and to evaluate the simulation results (that is, global elements along wall length, where each element consisted of
load-deformation responses, local shear deformation and five concrete fibers and four steel fibers, with finer discreti-
strain responses, and the observed failure mechanisms) were zation of the fibers located in the wall boundary. For FWE,
available in the referenced papers. Main characteristics of the the cross section was subdivided into many triangular
wall specimens are summarized in Table 2. fibers (for example, 1566 for Specimen SP4) as shown in
Fig. 10(b), with four types of fibers: 1) steel (one fiber for
MODEL GENERATION each bar); 2) concrete cover; 3) unconfined concrete; and 4)
Procedures used to generate analytical models of wall spec- confined concrete. For the SFI-MVLEM-FD, macro-fibers
imens selected are summarized in the following sections for in the boundary regions were narrower than web macro-
the five macroscopic modeling approaches described in the fibers, typically including two to four bars, to predict the

1644 ACI Structural Journal/November 2018


Table 2—Experimental data for planar wall test specimens
Specimen fc′, fyBE, ρvBE, ρh,web/ρv,web, Vmax/(Acv√fc′), Δu, Failure
ID Scale lw/t h/lw MPa MPa % % M/(Vlw) P/(Agfc′) MPa (psi) Vmax/Vn Mmax/Mn Δy, % % mode*
RW2 0.33 12 3.00 34.0 434 2.93 0.33 / 0.33 3.13 0.09 0.22 (2.7) 0.52 1.16 0.55 2.35 CB
R2 0.33 19 2.34 46.4 450 4.00 0.31 / 0.25 2.40 0.00 0.17 (2.1) 0.42 1.23 0.34 2.89 CB/LI
WSH6 0.49 13 2.02 45.6 576 1.54 0.25 / 0.46 2.26 0.11 0.30 (3.6) 0.83 1.11 0.31 2.04 CB/LI
RW-A15-
0.50 8 1.50 55.8 477 6.06 0.73 / 0.73 1.50 0.10 0.65 (7.8) 0.89 1.19 0.63 1.07 DC
P10-S78
S6 0.37 21 1.26 27.8 482 5.60 0.55 / 0.55 1.60 0.05 0.53 (6.4) 0.80 1.12 0.32 0.65 CB/LI
*
CB is concrete crushing/reinforcing bar buckling; LI is lateral instability of wall boundary; and DC is diagonal compression.
Note: 1 MPa = 145 psi.

Fig. 10—Vertical and horizontal model discretization of Specimen SP4: (a) P3D-SW; (b) FWE; (c) SFI-MVLEM-FD; (d)
SFI-MVLEM-SS; and (e) NLTM. (Note: Units are in.; 1 in. = 25.4 mm.)
flexural response of the wall more precisely (Fig. 10(c)), ness, and reloading towards the same strain and stress from
whereas length of the web segments was defined based on which unloading initiated.
the distances between the longitudinal web reinforcement The concrete model used in the FWE was calibrated
(two to four bars per macro-fiber). A similar yet simplified using an ascending parabola followed by a linear softening
approach for cross-sectional discretization was employed branch in compression (Scott et al. 1982), whereas tensile
in the SFI-MVLEM-SS and NLTM models, where the two strength was neglected, which corresponds to the well-
outer panels represent the confined wall boundaries (only known Concrete01 material model in OpenSees. Although
one per boundary) and the remaining equal-width panels Concrete01 assumes zero tension capacity of concrete,
represented the wall web (Fig. 10(d) and (e)). previous studies with FWE (Vásquez et al. 2016) showed
that this approach provides reasonable predictions of hyster-
Calibration of material models etic behavior of RC walls. For concrete cover and unconfined
Calibration of material parameters, which define the concrete, slope of the descending branch was determined
flexural and shear response characteristics of the wall using the concrete crushing energy proposed by Nakamura
models based on either material strain-stress or constitutive and Higai (2001), and a residual strength of zero was speci-
force-deformation behavior, is described in the following fied. Parameters of the constitutive relationship for confined
sections. concrete were calibrated according to model proposed by
Models based on material strain-stress behavior— Scott et al. (1982), with a residual strength corresponding to
Concrete stress-strain relationship: In the P3D-SW 20% of the peak strength.
model, calibration of the unconfined concrete stress-strain In the SFI-MVLEM-SS model, the monotonic envelope of
relationship was based on as-tested material properties the stress-strain model for unconfined concrete in compres-
reported, whereas the confined concrete relationship was sion proposed by Chang and Mander (1994) (OpenSees
calibrated to follow the Saatcioglu and Razvi (1992) model. material ConcreteCM; Kolozvari et al. 2018) was calibrated
For both cases, trilinear relationships with strength loss to agree with material properties obtained from cylinder tests
(Fig. 1(c)) were used to approximately represent concrete on day of testing (when available) by matching the compres-
strain-stress behavior; with an initial slope of Ec up to a sive strength, the strain at compressive strength, initial tangent
stress of 0.4fc′, a second linear segment up to maximum modulus, and the shape parameter defining the monotonic
stress of fc′, a plateau at fc′, and a linear strength loss segment stress-strain curve proposed by Tsai (1988). The post-peak
following the plateau. This monotonic trilinear relationship slope of the strain-stress curve was calibrated to match the
forms the envelope for the cyclic strain-stress behavior, monotonic envelope proposed by Saatcioglu and Razvi
characterized by unloading parallel to the initial elastic stiff- (1992). The stress-strain envelopes for confined concrete

ACI Structural Journal/November 2018 1645


were obtained using the confinement model by Mander et al. of reinforcing bars including yield strength and strain hard-
(1988). The tensile strength of concrete was determined from ening ratio. The tensile yield strength and strain-hardening
the relationship ft = 0.31√fc′ (MPa), and a value of 0.00008 parameters were modified according to empirical relations
was selected for the strain at the peak monotonic tensile stress proposed by Belarbi and Hsu (1994) to include the effect of
εt, as suggested by Belarbi and Hsu (1994). To represent tension stiffening on steel bars embedded in concrete. The
tension stiffening effects on concrete, post-crack shape of the parameters describing the cyclic stiffness degradation char-
monotonic tension envelope of the Chang and Mander (1994) acteristics of the reinforcing bars were calibrated as R0 = 20,
model was calibrated against the stress-strain relationship a1 = 0.925, and a2 = 0.15, as proposed originally by Mene-
proposed by Belarbi and Hsu (1994). gotto and Pinto (1973). Reinforcement buckling or fracture
In the NLTM, concrete in vertical and horizontal elements is not represented in the SteelMPF model.
was modeled as either confined or unconfined, depending NLTM uses the Giuffré-Menegoto-Pinto model (Filippou
on their location within the cross section. Concrete behavior et al. 1983), available in OpenSees as uniaxial material
was defined using the Yassin (1994) model implemented Steel02, to simulate the behavior of longitudinal reinforcing
in OpenSees as Concrete02. Concrete tensile behavior was bars. Bar buckling and/or rupture is not accounted for by
implemented for the vertical elements, based on model by the model. Yield strength and initial elastic modulus were
Scott et al. (1982), while tensile capacity of the horizontal calibrated to match the as-tested material properties, a strain
elements was set to zero. Concrete compressive strength was hardening ratio of 0.01 was used for all specimens, and
calibrated to the value reported for each wall specimen at the cyclic response parameters of R0 = 20, CR1 = 0.925, CR2 =
day of testing, and the linear post-peak (softening) branch 0.15 were used in the calibration, as recommend by Filippou
of the concrete model was regularized as per recommen- et al. (1983) and Lu and Panagiotou (2014).
dations by Coleman and Spacone (2001). The behavior of Models based on constitutive force-deformation
concrete diagonals was modeled using a parabolic compres- behavior—Calibration of the force-deformation relation-
sive stress-strain relationship up to fc′ based on the Fujii ships used in the SFI-MVLEM-FD formulation is presented
concrete model (refer to Hoshikuma et al. 1997), where the in the following paragraphs.
softening branches were regularized automatically since the Properties and modeling of vertical springs: In walls
material model receives feedback from the geometry of each with heavy boundary reinforcement, the boundary segments
diagonal. The tensile stress-strain relationship of concrete were typically modeled by combining two materials using the
diagonals was assumed linear up to the tensile strength ft, Parallel material available in OpenSees: 1) a material used
after which the concrete softens in an exponential manner to define the response of the concrete part of the segment,
according to Stevens et al. (1991). Further details on the and 2) a material used to define the response of the longitu-
behavior of concrete diagonals can be found in Lu and dinal bars. For example, in specimens SP4 and S6, material
Panagiotou (2014). VertSpringType2 developed by Kante (2005) was used to
Steel stress-strain relationship: In the P3D-SW model, model the concrete part of the segment (material model is not
the reinforcing steel stress-strain relationship implemented in publicly available), while the Steel02 (Menegotto and Pinto
Perform 3D was based on trilinear idealization of the actual 1973) or Hysteretic model (OpenSees, McKenna et al. 2000)
stress-strain relations obtained from reported material charac- was used to model longitudinal bars. When modeling the
terization tests. Symmetric behavior in tension and compres- concrete part, properties of confined concrete were consid-
sion was assumed. The stress-strain envelope first follows a ered as per the model by Mander et al. (1988). For simula-
linear segment with slope Es up to the yield stress fy, followed tion of lightly reinforced boundary regions and/or wall web
by a plateau, and a strain hardening up to the ultimate strength (segments between boundary regions), only one material
fu. Stiffness degradation under cyclic loading is accounted (for example, VertSpringType2) was used, combining the
for by using the energy factor defined in Perform 3D equal properties of concrete and the longitudinal bars.
to 0.75, indicating that the unloading and reloading stiffness Modeling of shear behavior—
values in the hysteretic stress-strain relationship were adjusted Models without shear-flexure interaction: In models
so that the area of the loop with stiffness degradation equals to with uncoupled axial/flexural and shear responses, shear
75% of the loop area without stiffness degradation. behavior is typically accounted for by using a horizontal
Material model parameters for the steel fibers in the FWE spring with specified force-deformation or strain-stress
were calibrated to match the material tests provided for each (backbone) relation. A relatively wide range of effective shear
wall specimen. A piecewise linear function is used to define stiffness recommendations for walls can be found in current
the backbone curve and the cyclic behavior follows Mene- provisions (ASCE 41-17 2017; LATBSDC 2017; PEER/
gotto and Pinto (1973) model. For the descending branch ATC 72 2010), as well as in the literature (Tran and Wallace
of the steel in compression, the buckling model proposed 2015; Gogus 2010), and it has been shown that prediction of
by Dhakal and Maekawa (2002a) was used, where the wall response can be significantly influenced by the choice
descending branch depends on the longitudinal bar diameter, of a shear backbone relation and an effective shear stiffness
its strength, and the stirrup spacing. value (Kolozvari and Wallace 2016). Uncoupled analytical
SFI-MVLEM-SS uses the uniaxial material model for models used in this study employ the following modeling
reinforcing steel proposed by Menegotto and Pinto (1973), approaches for simulating shear behavior:
available in OpenSees as SteelMPF (Kolozvari et al. 2018), 1. P3D-SW uses effective shear stiffness values recom-
which was calibrated to represent as tested material properties mended by PEER/ATC-72 (2010) Geff = 0.05Gc ≈ 0.02Ec for

1646 ACI Structural Journal/November 2018


Specimen SP4, while a higher value (Geff = 0.10Gc ≈ 0.04Ec) followed by a comparison of the accuracy of the models
was used for Specimen RW2, as the shear stress at devel- to predict key force and stiffness parameters derived from
opment of the nominal flexural capacity in this specimen the envelopes of the hysteretic load-deformation behavior
is significantly lower (0.17√fc MPa versus 0.65√fc′ MPa (“Envelope responses”).
[2.1√fc psi versus 7.8√fc′ psi]). A similar approach is used for Hysteretic behavior—Comparisons of the experimen-
other specimens considered herein. tally measured and analytically predicted hysteretic lateral
2. MVLEM-FD uses ShearSlip material, implemented in load versus total top displacement responses for the five RC
the in-house version of OpenSees (Kante 2005), to define wall specimens selected, obtained using the five different
the behavior of the shear spring within the model element modeling approaches, are presented in Fig. 11.
used for simulation of Specimen RW2. Because the response Results obtained using P3D-SW (Fig. 11(a.1) through
of this specimen was dominated by flexural deformations, a (a.5)), in which shear behavior is represented with a linear
simple linear-elastic model for shear response with effective elastic relationship, are in good agreement for specimens
shear stiffness of GAeff = 0.2EcAw was adopted. with flexure-dominated behavior, such as Specimens RW2,
3. FWE uses the Modified Compression Field Theory R2, and WSH6 (Fig. 11(a.1), (a.2), and (a.4)), where the
(Vecchio and Collins 1986) to define the shear backbone overall wall capacity is predicted within 5% of experimen-
behavior as described in the section “Distributed plasticity tally measure values. However, using the same modeling
fiber-based beam-column force wall element”. approach for Specimens SP4 and S6 (Fig. 11(a.2) and (a.5)),
Models with shear-flexure interaction: In analytical where nonlinear shear deformations are significant and
models that capture shear-flexural interaction (SFI-MV- shear-flexural interaction impacts the wall behavior, over-
LEM-FD, SFI-MVLEM-SS, and NLTM), shear stiffness and estimates measured wall capacity by approximately 31%
shear force-deformation behavior of each model element are and 15%, respectively. In addition, hysteretic properties of
incorporated directly in the element formulations and evolve the P3D-SW response are characterized with the absence of
according to the physical constitutive behavior assumed cyclic strength degradation, due to the simplified material
(versus use of a predefined shear force versus deformation models implemented in Perform 3D, and pinching character-
backbone relationship). Refer to the section “Description of istics of the wall specimens are either overestimated (Spec-
Macroscopic Modeling Approaches” for description of rele- imens RW2 and WSH6) or underestimated (Specimen R2)
vant models and modeling parameters used for shear behavior. by the model.
Load application—Application of lateral and axial loads Analytical predictions of wall behavior using the fiber-
on the wall models was performed in a similar manner for based FWE (Fig. 11(b.1) through (b.5)) are in good agree-
all models considered. First, a constant vertical force was ment with measured load-displacement responses, where
applied at the top of each wall model using force-controlled wall lateral load capacity is predicted within 10% of the
nonlinear analysis, followed by a displacement-controlled experimentally obtained wall capacity for all specimens
Newton-Raphson nonlinear analysis solution strategy that considered. For specimen WSH6 (Fig. 11(b.4)), degradation
imposes lateral displacements at the top of the wall as per the of wall capacity during the last loading cycle is also over-
loading protocols reported for each wall specimen, under the estimated by the model. Because this modeling approach
appropriate load pattern. For all walls analyzed, the lateral considers a regularized descending branch for the shear
load pattern was in the form of one horizontal force applied constitutive relationship at each section, the capacity of walls
at the top of the wall specimen, except for Specimen S6, with even significant shear deformations is predicted with
which was subjected to a proportional lateral load pattern good accuracy. Cyclic behavior is generally well predicted,
consisting of three horizontal forces applied at each story excluding overestimation of the area under the hysteretic
level (1.22, 2.13, and 3.05 m [48, 84, and 120 in.] from the loops for Specimen R2, as shown in Fig. 11(b.3).
wall bottom), and a bending moment applied at the top of the The analytical prediction of the global response predicted
test specimen, as described by Vallenas et al. (1979). using the MVLEM-FD is good for the slender RC wall
Specimen RW2 (Fig. 11(c.1)), which is characterized with
COMPARISON OF ANALYTICAL AND flexure-dominated behavior, where wall yield and ultimate
EXPERIMENTAL RESULTS capacity are predicted within 5% of the experimentally
Effectiveness of the analytical models considered in this measured values. In addition, SFI-MVLEM-FD accu-
study was evaluated based on the accuracy with which rately captures well the hysteretic behavior and ultimate
various experimentally measured response quantities are wall capacity (within 3%) of the remaining specimens
simulated. Response quantities evaluated include total, (Fig. 11(c.2) through (c.5)), even for cases where the spec-
flexural, and shear load-deformation responses and strains imen behavior was characterized with significant amount
within the plastic hinge region. of nonlinear shear deformations (SP4 and S6). Pinching
behavior is well respresented for all specimens. During the
Load-displacement responses final loading cycles of Specimens SP4 and S6, global buck-
The following two subsections provide discussion on the ling of the longitudinal bars in the boundary regions of both
accuracy of considered RC wall models to capture cyclic walls was observed during the experiments. For Specimen
load-deformation behavior of the wall specimens selected SP4, the central part of the wall between boundary regions
for the study. Predicted hysteretic response of walls is subsequently failed due to shear sliding causing considerable
first described for each specimen (“Hysteretic behavior”), reduction of the shear and overall strength of the wall. The

ACI Structural Journal/November 2018 1647


Fig. 11—Load-displacement responses for five wall specimens and five models considered. (Note: 1 mm = 0.0394 in.)
analysis results obtained using SFI-MVLEM-FD provided tion, plastic (residual) displacements, and modest pinching
further insight into failure of Specimen SP4 showing that behavior are well represented in the model results for all wall
the shear strength degraded mainly due to the considerable specimens, except for Specimen R2 (Fig. 11(d.3)) where
deterioration of the dowel mechanism, provided mainly by pinching is notably underestimated, possibly due to zero
the heavy boundary reinforcement. axial load applied on the wall. Although the analytical model
The SFI-MVLEM-SS captures well the lateral load captures the initiation of strength degradation in the concrete
capacity of all specimens at most of the applied drift levels stress-strain behavior within the boundary elements, the
(Fig. 11(d.1) through (d.5)) where analytically predicted significant strength loss observed in some of the tests is not
lateral load capacities are within the 2 to 8% range of the represented in the analysis results, because of the inability
experimentally measured wall capacities. In addition, cyclic of the model to simulate failure mechanisms associated
characteristics of the response, including stiffness degrada- with buckling or fracture of reinforcing bars or sliding shear

1648 ACI Structural Journal/November 2018


failure near the base of the walls (Tran and Wallace 2015). total, flexural, and shear deformations are also calculated for
Therefore, for walls experiencing such failure modes, the the analytical and experimental results, up to the cracking
SFI-MVLEM-SS model generally fails to capture strength point (that is, initial total KC,tot, flexural KC,F, and shear KC,Sh
loss at ultimate drift levels, resulting in overestimation of stiffness) and the yield point (that is, yield total KY,tot, flex-
their deformation (drift) capacity. ural KY,F, and shear KY,Sh stiffness).
Comparison between analytical results obtained using Comparison of the aforementioned experimental and
NLTM with experimental data presented in Fig. 11(d.1) analytical envelope response quantities is presented in
through (d.5) show that the mathematical model global Table A.1 of Appendix A,* whereas graphical comparisons
responses are in good agreement with experimental results, of the envelope strength quantities and effective stiffness
although generally larger uncracked stiffness is predicted values are illustrated in Fig. 13 and 14, respectively. It
by the model (refer to section “Envelope responses”) due should be noted that the average of positive and negative
to overlapping areas of diagonal concrete struts. After values is used for the comparisons, given that wall geometry
cracking and yielding occurs, the model under predicts the and loading conditions are symmetric. It can be concluded
post-yielding strength by up to 10%, except for Specimen from results presented in Fig. 13(a) and (c) that yield
S6 where the strength is overestimated by 8%, which is capacity (VY) and ultimate capacity (VU) are predicted with
associated with the shorter arm of the axial forces carried 5% accuracy for specimens with flexure-dominated behavior
by the extreme elements per discretization process used in (RW2, R2, and WSH6), whereas for Specimens SP4 and S6,
this study (refer to Fig. 10(e)). The general shape of the in which shear deformations considerably influence wall
hysteretic loops represented shows good agreement for behavior, the predicted yield and ultimate capacities are
all specimens, except for cycles after the onset of failure. predicted with 15% accuracy, compared to experimentally
The NLTM does not capture well the onset of strength loss, obtained values. The biggest overestimation of wall ultimate
which is mainly controlled by the strut strength degradation capacity for these two specimens can be observed for the
model defined by the β-curve (Fig. 9(b)). Although strength P3D-SW model, because this model does not incorporate
degradation observed in Specimens SP4, R2, and S6 can be shear-flexural interaction and assumes linear elastic shear
better captured by calibrating the β-curve parameters inde- behavior. Model predictions are considerably less accurate
pendently for each specimen, the same parameters of the in terms of displacements corresponding to wall yielding
β-curve were selected for all specimens, where the parame- (DY; Fig. 13(b)) and ultimate (DU; Fig. 13(d)) capacity. Like
ters were defined such that the strength of the struts degrade predicted capacities, the bigger scatter of results can be
at a low rate for increasing normal strains. For example, observed for Specimens SP4 and S6 because of the signif-
the β-curve parameters {εint, βint} and {εres, βres} shown in icant contribution of nonlinear shear deformations experi-
Fig. 9(b) were set to {0.04, 0.16} and {0.05, 0.09}, respec- enced by these specimens (refer to the section “Shear and
tively, which correspond to points that lie on the Vecchio flexural deformations” for shear deformations for Specimen
and Collins (1993) strength degradation curve. Prediction of SP4). Finally, most of the predicted lateral load values corre-
onset and progression of strength degradation observed in sponding to the maximum lateral displacement applied (VX;
Specimens SP4, R2, and S6 can be improved by calibrating Fig. 13(e)) are within an approximate 20% margin of error
the parameters of the β-curve for each specimen. The way compared to the experimental values obtained for most of the
the model was implemented in this study does not account wall specimens. However, there is significant scatter from
for bar buckling and/or rupture, nor the subsequent loss of one model to the next and from one specimen to the other,
strength of confined concrete. due to challenges in simulating the experimentally observed
Envelope responses—To quantify the ability of the models failure modes. The biggest dispersion can be observed for
to capture specimen strength and stiffness at various levels Specimen R2, where all models expect SFI-MVLEM-FD
of deformation, and to allow consistent comparison between tend to overestimate the lateral load on the wall at maximum
the load-deformation responses predicted by the modeling displacement by approximately 50 to 75% (that is, do not
approaches considered, envelopes of the hysteretic load- capture strength degradation).
deformation responses are generated from both analytical Figure 14(a) further reveals that overall initial stiffness of
and experimental data. The load-deformation envelope is the walls is generally overestimated by most models, as much
generated using four points on the measured/predicted load- as two times. P3D-SW tends to underestimate initial stiffness,
deformation curves, corresponding to: 1) cracking (C); 2) which is related to the low effective shear stiffness adopted in
yielding (Y); 3) ultimate (maximum) capacity (U); and 4) this study (that is, GAeff = GAg/20), while SFI-MVLEM-FD
maximum applied displacement (X), as illustrated in Fig. 12 provides the best stiffness predictions. Similar trends can be
for the positive loading direction of Specimen SP4. For observed among the other model results for all specimens.
Specimens RW2 and SP4, load-deformation envelopes were Note that overestimation of wall initial stiffness is primarily
considered for total (Fig. 12(a)), flexural (Fig. 12(b)), and due the inability of the models to account for micro-cracking
shear (Fig.  12(c)) deformations (flexural and shear hyster- in concrete (Palermo 2002) and strain penetration effects that
etic responses are presented in the section “Shear and flex- are not explicitly incorporated in the modeling approaches;
ural deformations”), whereas for the remaining specimens, note that micro-cracking does not affect FWE results as the
only load versus total lateral top displacement is considered. model ignores concrete tension capacity. Results presented in
Based on forces and deformations defining the envelope Fig. 14(c) and (d) suggest that the modeling approaches gener-
curves, effective secant stiffness values corresponding to ally overestimate initial flexural stiffness, whereas initial shear

ACI Structural Journal/November 2018 1649


Fig. 12—Load-deformation envelopes (positive loading only): (a) total; (b) flexural; and (c) shear.

Fig. 13—Comparison of load-deformation envelope points.


stiffness is generally underestimated for Specimen RW2 (flex- be noted that shear and flexural deformations of Specimen
ure-dominated wall), while significant scatter of results can be RW2 were measured only along the bottom 1828 mm (72 in.)
observed for the initial shear stiffness prediction for Specimen (that is, the first and second story) during the experiment.
SP4. Secant stiffness values to yield point corresponding to total Models with uncoupled shear and flexure behavior—The
displacement are generally predicted more accurately by all of fiber-based macroscopic models considered with uncou-
the analytical models, where the ratio of predicted to measured pled shear and flexural responses (P3D-SW, FWE, and
stiffness varies between 0.5 and 1.5, as illustrated in Fig. 14(b). MVLEM-FD) require specification of a shear force-defor-
However, Fig. 14(e) and (f) reveal that there is still consider- mation backbone (for example, linear-elastic, elasto-plastic),
able scatter among model predictions for the secant flexural where the predicted shear behavior of the wall is influenced
and shear stiffness to yield point, where the ratio of predicted to by the choice of the model parameters specified. Uncou-
measured stiffness ranges between 0.5 and 2.0, with the excep- pled analytical models considered in this study incorporate
tion of shear stiffness to yield point for Specimen RW2, which linear-elastic shear behavior, where the effective shear stiff-
is overestimated four times by the FWE. ness for each model is specified as described in the section
“Modeling of shear behavior”.
Shear and flexural deformations Results presented in Fig. 15(a.2), (b.2), and (c.2) reveal
Comparisons of the experimentally measured and analyt- that nonlinear shear deformations over the plastic hinge
ically predicted lateral load versus flexural and shear defor- region (that is, bottom 914 mm [36 in.]) are significantly
mation components over the bottom 914 mm (36 in.) (that is, underestimated by the uncoupled models with linear-elastic
the first story), are presented for Specimen RW2 (3.0 aspect shear behavior and that this approach does not represent real-
ratio, slender wall, low shear stress) in Fig. 15, while lateral istically the hysteretic shear behavior observed in the tests.
load versus flexural and shear deformations corresponding to Because nonlinear shear deformations constituted approxi-
the top of the wall are compared for Specimen SP4 (1.5 aspect mately 30% of total lateral displacement over the first story
ratio, medium-rise wall, high shear stress) in Fig. 16. It should of Specimen RW2, uncoupled modeling approaches tend

1650 ACI Structural Journal/November 2018


Fig. 14—Comparison of effective stiffness values.
to overestimate flexural deformations by approximately 10 RW2, respectively, which are load levels that are below their
to 20% within this region (Fig. 15(a.1), (b.1), and (c.1)). nominal shear capacities.
However, the assumption of linear elastic shear force- The SFI-MVLEM-FD model was able to represent the
deformation relationship does not have significant impact on significant shear degradation during the last loading cycle of
overall wall load versus top deformation behavior (Fig. 11) the test on Specimen SP4, which was mainly associated in
because overall wall behavior was flexure dominated with the model results with degradation of the dowel mechanism
shear deformations contributing to the total top lateral and sliding shear at the bottom of the wall (Fig. 16(c.2)); note
displacement by approxiately 10% only. that failure mechanism observed during the exprement (Tran
In contrast, the behavior of Specimen SP4 was signifi- and Wallace 2015) was initiated with concrete crushing and
cantly influenced by nonlinear shear deformations, where buckling of reinforcement in the wall boundaries, which led
their contribution to total top displacement was approx- to shear sliding adjecent to the wall-foundation interface.
imately 40%. Therefore, using the assumption of linear- Shear deformation contribution to total top displacement
elastic shear behavior significantly underestimates nonlinear was predicted by the model to be 39% on average, whereas
shear deformations and overestimates nonlinear flexural 31% contribution was obtained in the test results. The contri-
deformations, as illustrated in Fig. 16(a) and (b). For all butions and importance of three shear mechanisms (aggre-
uncoupled models, using an inelastic constitutive model gate interlock HSA, dowel HSD and horizontal reinforce-
(backbone) for shear will likely improve shear deforma- ment HSS) were varied over the test. At the beginning of
tions predictions. However, two major issues related to this the test, the shear response was predominantly influenced
approach are problematic: 1) lateral load capacity of the by the aggregate interlock spring (HAS), whereas the impor-
wall will depend on the capacity assigned to adopted shear tance of the dowel (HSD) and the horizontal reinforcement
force-deformation behavior; and 2) after the shear capacity (HSS) mechanism was increased after the HSA mechanism
is reached, the majority of wall lateral displacements will be was deteriorated.
associated with nonlinear shear deformations (due to small The SFI-MVLEM-SS model captures well the magnitudes
post-yield shear stiffness), which will lead to significant of nonlinear flexural and shear deformation components for
underestimation of vertical strains in the wall (Kolozvari and Specimen SP4 (medium-rise wall, Fig. 16(d.1) and (d.2))
Wallace 2016). Therefore, it may not be possible to accu- throughout the cyclic loading history, except at the ultimate
rately simulate, using a mechanical/behavioral approach, loading cycle where the model underestimates the large shear
the global and local response characteristics of walls expe- deformations that occured during the test due to the shear sliding
riencing significant shear deformations and shear-flexural failure. Shear-resisting mechanisms currently incorporated in
interaction behavior using uncoupled modeling approaches. the SFI-MVLEM-SS do not consider strength degradation
Models with shear-flexure coupling—Results presented in due to detoriation of shear strength within a model element
Fig. 15 and 16 reveal that the analytical models that incor- (except crushing of concrete along diagonal directions), and
porate axial/flexural-shear interaction at the model element therefore, strength loss associated with failure mechanisms
level (SFI-MVLEM-FD, SFI-MVLEM-SS, and NLTM) can such as cyclic degradation in dowel action, cyclic degrada-
capture nonlinear flexural and shear deformations, as well as tion in shear aggregate interlock behavior, or reinforcement
their coupling throughout the entire cyclic loading history. buckling is not represented in the model results. For Specimen
As revealed in both experimental and analytical results, RW2 (slender wall), the model underestimates the nonlinear
flexural and shear yielding occur almost simultaneously for shear deformations measured over the first story height of the
both wall specimens at a lateral load level of approximately wall by approximately 50% (Fig. 15(d.2)). However, given
800 and 380 kN (180 and 25 kip) for Specimens SP4 and the flexure-dominated behavior of this specimen, and the

ACI Structural Journal/November 2018 1651


Fig. 15—Flexural and shear load-deformation responses for Specimen RW2. (Note: 1 mm = 0.0394 in.; 1 kN = 0.2248 kip.)

Fig. 16—Flexural and shear load-deformation responses for Specimen RW-A15-P10-S76. (Note: 1 mm = 0.0394 in.; 1 kN =
0.2248 kip.)
small magnitudes of the shear deformations measured, under- Local responses
estimation of shear deformations by SFI-MVLEM-SS model Figure 17 depicts comparison of analytically predicted
for this specimen did not affect the prediction of the overall and experimentally obtained vertical (longitudinal) strain
specimen hysteretic behavior. profiles along the wall base for Specimens RW2 and SP4,
A good agreement between the experimental results and the corresponding to drift levels of 0.5, 1.0, and 2.0%. The
response simulated by the NLTM is observed for both spec- experimental strains are measured using vertical LVDTs
imens in terms of relative contribution of shear and flexural over a gauge length of 230 and 355 mm (9 and 14 in.) for
deformations to total lateral displacement, as well as their Specimens SP4 and RW2, respectively, whereas the analyt-
hysteretic response shape. For Specimen RW2 (Fig. 15(e.1) ical strain predictions are obtained from the bottommost
and (e.2)), the model overpredicts shear deformations by element of each wall model considered, according to model
22% on average comparing to experimental data, whereas discretization described in the section “Mesh information”.
for Specimen SP4 (Fig. 16(e.1) and (e.2)), the ultimate shear All of the macroscopic modeling approaches considered
displacement after strength loss is underpredicted by 16% in this study, except the NLTM, employ the plane-sections-
in the positive direction and 40% in the negative direction remain-plane kinematic assumption. Therefore, analytically
(29% in average). obtained strain profiles for the plane-section models can
be defined with only two points at wall ends, with a linear
distribution of vertical strains in between. However, as can
also be observed in Fig. 17, experimentally measured strain

1652 ACI Structural Journal/November 2018


Fig. 17—Vertical strain profiles at various drift levels: (a) RW2; and (b) SP4. (Note: 1 mm = 0.0394 in.)
profiles are typically nonlinear, particularly at intermediate tions in terms of stiffness and strength. Similarly, predic-
and large drift levels. Based on comparison of the results, it tion of compression strains could have marginal impact on
can be deduced that at low-to-intermediate drift levels (that wall initial stiffness, which also depends on material model
is, 0.5% and 1.0%) maximum tensile strains are predicted for concrete used. However, accurate prediction of strains
with approximately 20% accuracy for Specimen RW2 is important for prediction of wall deformation capacity,
(Fig. 17(a.1)) and overestimated up to 2.0 times at the wall as degradation of wall capacity is typically associated with
boundary for Specimen SP4 (Fig. 17(b.1)) by most of the concrete crushing, buckling of reinforcing bars, out-of-plane
models. Only NLTM underestimates tension strains by 50% instability, and other mechanisms that are dependent on
at 1.0% drift of Specimen RW2, and only SFI-MVLEM-FD the strain history experienced within the wall plastic hinge
underestimates compression strains by 70% at 1.0% drift region. Therefore, accurate estimation of strain responses
of Specimen SP4. At same drift levels of 0.5% and 1.0%, is an important aspect of performance-based seismic
compression strains are underestimated by approximately design, and given the widespread use of macroscopic
30 to 40% for all models except SFI-MVLEM-FD, which models, analysis results for strains should be interpreted
predicts compression strains accurately for both specimens with a certain amount of conservatism, especially when
considered except at a drift of 1.0% for Specimen SP4 plane-section models are used.
(Fig. 17(b.2)). At a drift level of 2.0%, where both wall
specimens experienced significant nonlinear deformations, SUMMARY AND CONCLUSIONS
SFI-MVLEM-SS and FWE models overestimate maximum This paper summarizes results of comparative studies
tensile strains by approximately 10 to 30% for both wall conducted using five conceptually different macroscopic
specimens, whereas P3D-SW and SFI-MVLEM-FD models analytical models for RC walls available in the litera-
overestimate tensile strains by as much as 100%, particu- ture. Modeling approaches selected represent state-of-
larly for Specimen SP4 (medium-rise wall, Fig. 17(b.3)). the-art and state-of-the-practice in nonlinear macroscopic
Regarding compressive strains at a 2.0% drift, the SFI-MV- modeling of RC structural walls, and include: 1) shear wall
LEM-FD model predictions agree well with test results element available in Perform 3D (Computers and Struc-
for both specimens, P3D-SW and SFI-MVLEM-SS model tures Inc. 2005); 2) fiber-based nonlinear beam-column
predictions agree well with compression strains measured force wall element (Vásquez et al. 2016); 3) Multiple-Ver-
on the medium-rise wall Specimen SP4 (Fig. 17(b.3)), while tical-Line-Element-Model with uncoupled and coupled
FWE for overestimates compressive strains of this specimen axial/flexural and shear responses based on force-defor-
by a factor of 2.0 (Fig. 17(b.3)). For slender Specimen RW2 mation material behavior (Fischinger et al. 2012, 2017); 4)
(Fig. 17(a.3)), compressive strains predicted by all models, Multiple-Vertical-Line-Element-Model with coupled axial/
except SFI-MVLEM-FD, are only approximately 30 to 50% flexural and shear responses based on strain-stress material
of the experimentally measured strains. Finally, the NLTM behavior (Kolozvari et al. 2015a,b); and 5) nonlinear truss
can capture the observed nonlinear strain distributions along element (Panagiotou et al. 2012). The modeling approaches
the length of the wall specimens, which leads to generally considered are based on either two-node beam-column
improved predictions of the strain profiles, and yet compres- element formulations (Models 2 through 4), or four-node
sive and tensile strains measured in Specimen RW2 are still macroscopic element formulations (Models 1 and 5). In
underestimated by the NLTM (Fig. 17(a)). modeling approaches 1 through 4, the wall cross section is
The observed discrepancy in prediction of tensile strains discretized into a number of material fibers, the behavior of
has only modest influence on the global response predic- which is represented with either uniaxial material behavior

ACI Structural Journal/November 2018 1653


(models with uncoupled P/M-V behavior) or biaxial concrete ences can be observed between models that couple shear and
behavior with additional shear-resisting mechanisms (models flexural behavior and models that are based on uncoupled
with coupled P/M-V behavior), and the plane-sections-re- behavior. The following conclusions can be drawn regarding
main-plane assumption. Modeling approach 5 is based on the capability of the uncoupled and coupled macroscopic
a four-node macroscopic element defined using uniaxial models considered, to simulate nonlinear flexural and shear
sub-elements in horizontal, vertical, and diagonal directions, load-deformation responses of walls:
which allows nonlinear distribution of strains along the 1. Models with uncoupled P/M-V responses:
wall cross section. Analytical results were compared with a. Assumption of linear elastic shear behavior provides
experimental data obtained for five planar RC wall speci- estimations of shear deformations at wall yielding
mens with rectangular cross sections subjected to cyclic that are between 0.5 and 4.0 times the experimentally
uni-directional loading, characterized with a range of wall measured shear deformations, while maximum shear
properties including aspect ratio (moderately slender and deformations are only about 5% of the maximum
slender walls), axial load (0, 5, and 10% of wall axial load shear deformations measured in the experiments due to
capacity), reinforcement detailing, and failure mode (flex- inability of the modeling approach to capture nonlinear
ural or shear-flexural). Brief descriptions of the analytical shear behavior.
models are provided, followed by information on calibration b. Effective stiffness of the linear-elastic shear behavior
of the model geometry and material constitutive relation- directly influences overall wall strength and stiffness
ships, as well as the analysis procedure used. Comparison and relative contributions of shear and flexural defor-
between experimental results and analytical predictions mations to total lateral displacement, which could lead
are conducted at global and local response levels including to significant bias in estimation of wall shear forces
lateral load versus total, flexural and shear lateral deforma- and lateral displacements in the response history anal-
tions (hysteretic responses and backbone curve parameters), yses. By manipulating this parameter, it is possible to
as well as vertical strain profiles along the base of the walls. obtain reasonable predictions of wall force-displace-
Based on the comprehensive comparisons between the ment responses, as well as levels of shear deformation.
analytical and experimental results, the following conclu- However, hysteretic nonlinear shear behavior cannot be
sions can be made regarding the accuracy and effectiveness captured when linear-elastic relation for shear is used,
of currently available macroscopic models for RC structural and the value of effective shear stiffness is typically
walls in terms of the overall load-deformation wall behavior: chosen in an ad-hoc manner.
1. Cracking and yielding strength of the walls is generally 2. Models with coupled P/M-V responses:
well predicted by all analytical models considered, where a. All coupled modeling approaches considered are
most of the analytical results are within ±10% of the range capable of capturing both nonlinear flexural and shear
of values estimated from the experiments; only for one deformations and their coupling throughout cyclic
modeling approach and one specimen the difference is 17%. loading history.
2. Ultimate lateral load capacity of wall specimens with b. The magnitudes and contributions of flexural and
flexure-dominated behavior is well estimated by all analyt- shear deformations to global responses are predicted
ical models considered. However, for specimens that with a 30% accuracy for most of the models, where
experienced significant shear deformations, models with highly pinched shear hysteresis and absence of pinching
uncoupled P-M-V behavior tend to overestimate maximum in flexural hysteresis is captured well.
wall load capacity by up to 30%, while coupled modeling c. The amount of nonlinear flexural and shear deformations
approaches capture wall capacity with ±10% deviation from may be modestly influenced by parameters of constitutive
the experimentally measured value. models incorporated to represent various shear resisting
3. Initial stiffness is generally not predicted accurately by mechanisms along concrete cracks (for example, aggre-
the models, and all models tend to overestimate pre-cracked gate interlock in concrete and dowel action).
stiffness of the walls by factors of approximately 2.0 to 4.0. Comparison between experimentally measured and
This is primarily due to inability of the models to account for analytically predicted vertical strain profiles at the base of
micro-cracking in concrete and strain penetration effects that the walls reveal that macroscopic models are generally not
are not explicitly incorporated in the modeling approaches. reliable for prediction of local strains. The following obser-
4. Secant yield stiffness (post-cracked stiffness) is predicted vations can be made based on results presented in this paper:
within a ±50% range of experimentally estimated values. 1. The plane-sections-remain-plane assumption incor-
5. The shape of hysteretic loops is generally well captured porated in most macroscopic model formulations provides
by all analytical models, although it is observed to be sensi- good approximation of strain distributions at low drift levels,
tive to properties of the material modeling parameters used. prior to yielding of wall boundary reinforcement.
For example, use of concrete material models that do not 2. At post-yield drift levels, macroscopic wall models may
consider gradual gap closure generally leads to an overesti- overestimate tensile strains by as much as a factor of 2.0.
mation of pinching in the load-deformation responses, while 3. Compressive strains are generally underestimated
application of zero axial load could lead to underestimation 2.0 to 3.0 times by the macroscopic modeling approaches
of pinching behavior by some models. considered, where larger discrepancy between analytical and
Based on comparison of measured and predicted flexural experimental compressive strains is observed for the flex-
and shear load-deformation responses, significant differ- ure-controlled (slender) wall specimen. Predictions obtained

1654 ACI Structural Journal/November 2018


using models that capture shear-flexure interaction are typi- research interests include advancing numerical modeling for earthquake
response and performance assessment of concrete buildings and bridges.
cally more accurate.
4. The discrepancy in prediction of tensile strains and Kutay Orakcal is an Associate Professor in the Department of Civil
compressive strains does not necessarily influence signifi- Engineering at Bogazici University, Istanbul, Turkey. He received his BS
from Middle East Technical University, Ankara, Turkey, and his MS and
cantly the global response predictions in terms of wall stiff- PhD from UCLA. His research interests include structural and earthquake
ness and strength, but it plays significant role in prediction engineering, with emphasis on nonlinear seismic response evaluation of
of wall deformation capacity, which is an important aspect reinforced concrete components and systems through analytical modeling,
laboratory testing, and field testing.
of performance-based seismic design. Therefore, analysis
results for strains should be interpreted with certain amount of Jorge A. Vásquez is a Structural Engineer at PUC, where he received
conservatism, especially when plane-section models are used. his MS in 2015. He then joined Centro Nacional de Investigación para la
Gestión Integrada de Desastres Naturales (CIGIDEN) at PUC, where he
Future research in modeling of RC wall systems could teaches and conducts research on nonlinear behavior of reinforced concrete
include response simulation of walls subjected to higher structures, and seismic risk and resilience of critical infrastructure.
axial loads, analysis of coupled wall systems, modeling
John W. Wallace, FACI, is a Professor of civil engineering at UCLA. He
and analysis of walls with nonrectangular cross sections is a member of ACI Committees 318, Structural Concrete Building Code;
subjected to bidirectional lateral loading, and simulation of 369, Seismic Repair and Rehabilitation; 374, Performance-Based Seismic
dynamic shake table tests. As well, future work can focus on Design of Concrete Buildings; and ACI Subcommittee 318-H, Seismic
Provisions. His research interests include response and design of buildings
addressing shortcomings of the current modeling approaches, and bridges to earthquake actions, laboratory and field-testing of structural
development and validation of novel modeling methodolo- components and systems, and seismic structural health monitoring.
gies, and development of a database of both existing and
new wall models available on web-based platforms (for ACKNOWLEDGMENTS
example, the Natural Hazards Engineering Research Infra- Funding for U.S. participants was provided by National Science Founda-
tion Grant CMMI-1446423: SAVI: Virtual International Institute for Perfor-
structure [NHERI] website, and the website of SAVI: Virtual mance Assessment of Wall Systems. Funding for participants outside of the
International Institute for Performance Assessment of Wall United States was provided by the individual countries, including Ministry
Systems). of Business, Innovation and Employment (MBIE, New Zealand), Chile,
Colombia, Turkey, and Slovenia. The authors would also like to thank A.
Martinez, an undergraduate student at Universidad Del Norte, Colombia,
AUTHOR BIOS for help with numerical analyses.
Kristijan Kolozvari is an Assistant Professor in the Department of Civil
and Environmental Engineering at California State University, Fullerton,
Fullerton, CA. He received his BS from the University of Belgrade, Belgrade, REFERENCES
Serbia, and his MS and PhD from the University of California, Los Angeles ASCE 41-17, 2017, “Seismic Evaluation and Retrofit of Existing Build-
(UCLA), Los Angeles, CA. His research interests include development and ings,” American Society of Civil Engineers, Reston, VA, 550 pp.
application of innovative analytical tools for nonlinear analysis of rein- Baker, C.; Lowes, N.; and Lehman, D. E., 2016, “Recommendations for
forced concrete structures, performance-based seismic design, seismic Modeling the Nonlinear Response of Slender Reinforced Concrete Walls
retrofit, tall building behavior and design, and earthquake resiliency. Using PERFORM-3D,” Proceedings of the 2016 Convention of Structural
Engineering Association of California (SEAOC), Maui, HI.
Carlos A. Arteta is a Professor of civil engineering at Universidad del Belarbi, A., and Hsu, T. C., 1994, “Constitutive Laws of Concrete in
Norte, Barranquilla, Colombia. He received his BS from Universidad del Tension and Reinforcing Bars Stiffened By Concrete,” ACI Structural
Norte and his MEng and PhD from the University of California, Berkeley, Journal, V. 91, No. 4, July-Aug., pp. 465-474.
Berkeley, CA. His research interests include the design and behavior of Chang, G. A., and Mander, J. B., 1994, “Seismic Energy Based Fatigue
concrete structures and earthquake engineering. Damage Analysis of Bridge Columns: Part I – Evaluation of Seismic
Capacity,” Technical Report No. NCEER-94-0006, National Center for
Matej Fischinger is a Professor of earthquake engineering and reinforced Earthquake Engineering Research (NCEER), State University of New
concrete structures at the University of Ljubljana (UL), Ljubljana, Slovenia. York, Buffalo, NY, 232 pp.
Chowdhury, S. R., and Orakcal, K., 2013, “Analytical Modeling of
Sofia Gavridou is a Lecturer at UCLA. She received her diploma in Columns with Inadequate Lap Splices,” ACI Structural Journal, V. 110,
civil engineering from Aristotle University of Thessaloniki, Thessaloniki. No. 5, Sept.-Oct., pp. 735-744.
Greece; her MSc from the Rose School, Pavia, Italy; and her PhD from Coleman, J., and Spacone, E., 2001, “Localization Issues in Force-Based
UCLA. Her research interests include design and analysis of precast Frame Elements,” Journal of Structural Engineering, ASCE, V. 127, No. 11,
concrete structures, large-scale structural testing, and nonlinear analysis pp. 1257-1265. doi: 10.1061/(ASCE)0733-9445(2001)127:11(1257)
and seismic design of tall reinforced concrete buildings. Computers and Structures Inc, 2005, “PERFORM-3D,” Berkeley, CA.
Dashti, F.; Dhakal, R. P.; and Pampanin, S., 2014, “Simulation of Out-of-
ACI member Matias A. Hube is a Structural Engineer from Pontifica Plane Instability in Rectangular RC Structural Walls,” Second European
Universidad Católica de Chile (PUC), Santiago, Chile, where he received Conference on Earthquake Engineering and Seismology, European Asso-
his MS in 2002. He received his PhD from the University of California, ciation of Earthquake Engineering (EAEE); European Seismological
Berkeley, in 2009. His research interests include reinforced concrete, earth- Commission (ESC), 2ECESS, Istanbul, Turkey.
quake engineering, and bridge engineering. Dazio, A.; Beyer, K.; and Bachmann, H., 2009, “Quasi-Static Cyclic Tests
and Plastic Hinge Analysis of RC Structural Walls,” Engineering Struc-
Tatjana Isaković is Professor and Head of the Chair of Structural and tures, V. 31, No. 7, pp. 1556-1571. doi: 10.1016/j.engstruct.2009.02.018
Earthquake Engineering at UL, where she received her PhD in 1996. Her Dhakal, R., and Maekawa, K., 2002a, “Path-Dependent Cyclic
research interests include seismic design of reinforced concrete structures Stress-Strain Relationship of Reinforcing Bar Including Buckling,”
(bridges, walls, and precast industrial buildings), seismic strengthening Engineering Structures, V. 24, No. 11, pp. 1383-1396. doi: 10.1016/
and retrofit, as well as seismic base isolation. S0141-0296(02)00080-9
Dhakal, R., and Maekawa, K., 2002b, “Reinforcement Stability and
Laura Lowes, FACI, is the Department Chair and Conner Professor of Fracture of Cover Concrete in Reinforced Concrete Members,” Journal
Civil and Environmental Engineering at the University of Washington, of Structural Engineering, ASCE, V. 128, No. 10, pp. 1253-1262. doi:
Seattle, WA. She is a member of ACI Committee 369, Seismic Repair and 10.1061/(ASCE)0733-9445(2002)128:10(1253)
Rehabilitation; Subcommittess 318-C, Safety, Serviceability, and Analysis, Dulacska, H., 1972, “Dowel Action of Reinforcement Crossing Cracks
and 318-N, Nonlinear Dynamic Analysis; and Joint ACI-ASCE Committee in Concrete,” ACI Journal Proceedings, V. 69, No. 12, Dec., pp. 754-757.
447, Finite Element Analysis of Reinforced Concrete Structures. Her Elwood, K. J., and Moehle, J. P., 2003, “Shake Table Tests and Analyt-
ical Studies on the Gravity Load Collapse of Reinforced Concrete Frames,”

ACI Structural Journal/November 2018 1655


PEER Report 2003/01, University of California, Berkeley, Berkeley, CA, quake Engineering Research Center, University of California, Berkeley,
346 pp. Berkeley, CA.
Filippou, F. C.; Popov, E. P.; and Bertero, V. V., 1983, “Effects of Bond Mander, J. B.; Priestley, M. J. N.; and Park, R., 1988, “Theoret-
Deterioration on Hysteretic Behavior of Reinforced Concrete Joints,” ical Stress-Strain Model for Confined Concrete,” Journal of Struc-
Report UCB/EERC-83/19, Earthquake Engineering Research Center, tural Engineering, ASCE, V. 114, No. 8, pp. 1804-1826. doi: 10.1061/
University of California, Berkeley, Berkeley, CA, 212 pp. (ASCE)0733-9445(1988)114:8(1804)
Fischinger, M.; Isaković, T.; and Kante, P., 2004, “Implementation of Massone, L. M.; Orakcal, K.; and Wallace, J. W., 2006, “Shear-Flexure
a Macro Model to Predict Seismic Response of RC Structural Walls,” Interaction for Structural Walls,” Deformation Capacity and Shear Strength
Computers and Concrete, V. 1, No. 2, pp. 211-226. doi: 10.12989/ of Reinforced Concrete Members under Cyclic Loading, SP-236, American
cac.2004.1.2.211 Concrete Institute, Farmington Hills, MI, pp. 127-150.
Fischinger, M.; Kante, P.; and Isaković, T., 2017, “Shake-Table Response McKenna, F.; Fenves, G. L.; Scott, M. H.; and Jeremic, B., 2000, “Open
of a Coupled RC Wall with Thin T-Shaped Piers,” Journal of Structural System for Earthquake Engineering Simulation (OpenSees),” Pacific Earth-
Engineering, ASCE, V. 143, No. 5, p. 04017004 doi: 10.1061/(ASCE) quake Engineering Research Center, University of California, Berkeley, CA.
ST.1943-541X.0001718 Menegotto, M., and Pinto, E., 1973, “Method of Analysis for Cyclically
Fischinger, M.; Rejec, K.; and Isaković, T., 2012, “Modeling Inelastic Loaded Reinforced Concrete Plane Frames Including Changes in Geometry
Shear Response of RC Walls,” Proceedings, 15th World Conference on and Non-Elastic Behavior of Elements under Combined Normal Force and
Earthquake Engineering, Lisbon, Portugal, No. 2120. Bending,” Proceedings, IABSE Symposium, Lisbon, Portugal.
Fischinger, M.; Rejec, K.; and Isaković, T., 2014, “Inelastic Shear Nakamura, H., and Higai, T., 2001, “Compressive Fracture Energy
Response of RC Walls: A Challenge in Performance Based Design and and Fracture Zone Length of Concrete,” Modeling Inelastic Behavior of
Assessment,” Performance-Based Seismic Engineering: Vision for an RC Structures under Seismic Loads, American Society of Civil Engineers,
Earthquake Resilient Society, M. Fischinger, ed., Springer, Dordrecht, Reston, VA, pp. 471-487.
Germany, pp. 347-363. NEES, 2016, “The NEES Databases,” Network for Earthquake Engi-
Fischinger, M.; Vidic, T.; Selih, J.; Fajfar, P.; Zhang, H. Y.; and neering Simulation, https://datacenterhub.org/resources/395. (last accessed
Damjanic, F. B., 1990, “Validation of a Macroscopic Model for Cyclic Aug. 7, 2018)
Response Prediction of RC Walls,” Computer Aided Analysis and Design of Neuenhofer, A., and Filippou, F. C., 1998, “Geometrically
Concrete Structures, N. B. Bicanic and H. Mang, eds., V. 2, Pineridge Press, Nonlinear Flexibility-Based Frame Finite Element,” Journal of Struc-
Swansea, UK, pp. 1131-1142. tural Engineering, ASCE, V. 124, No. 6, pp. 704-711. doi: 10.1061/
Gérin, M., and Adebar, P., 2009, “Simple Rational Model for (ASCE)0733-9445(1998)124:6(704)
Reinforced Concrete Subjected to Seismic Shear,” Journal of Struc- NHERI, 2018, “DesignSafe-CI,” Natural Hazards Engineering Research
tural Engineering, ASCE, V. 135, No. 7, pp. 753-761. doi: 10.1061/ Infrastructure, www.designsafe-ci.org. (last accessed Aug. 7, 2018)
(ASCE)0733-9445(2009)135:7(753) Oesterle, R.; Aristizabal-Ochoa, J.; Fiorato, A.; Russel, H.; and Corley,
Gogus, A., 2010, “Structural Wall Systems—Nonlinear Modeling and W., 1979, “Earthquake Resistant Structural Walls—Tests of Isolated Walls:
Collapse Assessment of Shear Walls and Slab-Column Frames,” PhD disser- Phase II,” Portland Cement Association, Skokie, IL, 327 pp.
tation, University of California Los Angeles, Los Angeles, CA, 243 pp. Oesterle, R. G.; Fiorato, A. E.; Johal, L. S.; Carpenter, J. E.; and Corley,
Gomes, A., and Appleton, J., 1997, “Nonlinear Cyclic Stress-Strain W. G., 1976, “Earthquake Resistant Structural Walls—Tests of Isolated
Relationship of Reinforcing Bars Including Buckling,” Engineering Struc- Walls,” Report to the National Science Foundation, Construction Tech-
tures, V. 19, No. 10, pp. 822-826. doi: 10.1016/S0141-0296(97)00166-1 nology Laboratories, Portland Cement Association, Skokie, IL, 315 pp.
Hoshikuma, J.; Kawashima, K.; Nagaya, K.; and Taylor, A. W., 1997, Orakcal, K.; Conte, J. P.; and Wallace, J. W., 2004, “Flexural Modeling
“Stress-Strain Model for Confined Reinforced Concrete in Bridge Piers,” of Reinforced Concrete Structural Walls—Model Attributes,” ACI Struc-
Journal of Structural Engineering, ASCE, V. 123, No. 5, pp. 624-633. doi: tural Journal, V. 101, No. 5, Sept.-Oct., pp. 688-698.
10.1061/(ASCE)0733-9445(1997)123:5(624) Orakcal, K.; Ulugtekin, D.; and Massone, L. M., 2012, “Constitutive
Jiang, H., and Kurama, Y., 2010, “Analytical Modeling of Medium-Rise Modeling of Reinforced Concrete Panel Behavior under Cyclic Loading,”
Reinforced Concrete Shear Walls,” ACI Structural Journal, V. 107, No. 4, Proceedings, 15th World Conference on Earthquake Engineering, Lisbon,
July-Aug., pp. 400-410. Portugal, 11 pp.
Kabeyasawa, T.; Shiohara, H.; Otani, S.; and Aoyama, H., 1983, Orakcal, K., and Wallace, J. W., 2006, “Flexural Modeling of Reinforced
“Analysis of the Full-Scale Seven-Story Reinforced Concrete Test Concrete Walls—Experimental Verification,” ACI Structural Journal,
Structure,” Journal of the Faculty of Engineering, The University of Tokyo, V. 103, No. 2, Mar.-Apr., pp. 196-206.
V. 37, No. 2, pp. 431-478. Palermo, D., 2002, “Behaviour and Analysis of Reinforced Concrete
Kante, P., 2005, “Seismic Vulnerability of RC Walls,” PhD dissertation, Walls Subjected to Reversed Cyclic Loading [Microform],” PhD disser-
University of Ljubljana, Ljubljana, Slovenia, 243 pp. tation, Department of Civil & Mineral Engineering, University of Toronto,
Kolozvari, K.; Orakcal, K.; and Wallace, J. W., 2015a, “Modeling of Toronto, ON, Canada, 372 pp.
Cyclic Shear-Flexure Interaction in Reinforced Concrete Structural Walls. Panagiotou, M.; Restrepo, J. I.; Schoettler, M.; and Kim, G., 2012,
Part I: Theory,” Journal of Structural Engineering, ASCE, V. 141, No. 5, “Nonlinear Cyclic Truss Model for Reinforced Concrete Walls,” ACI Struc-
p. 04014135 doi: 10.1061/(ASCE)ST.1943-541X.0001059 tural Journal, V. 109, No. 2, Mar.-Apr., pp. 205-214.
Kolozvari, K.; Tran, A. T.; Orakcal, K.; and Wallace, J. W., 2015b, PEER/ATC 72, 2010, “Modeling and Acceptance Criteria for Seismic
“Modeling of Cyclic Shear-Flexure Interaction in Reinforced Concrete Design and Analysis of Tall Buildings,” Applied Technology Council,
Structural Walls. Part II: Experimental Validation,” Journal of Structural Pacific Earthquake Engineering Research Center, Berkeley, CA, 147 pp.
Engineering, ASCE, V. 141, No. 5, p. 04014136 doi: 10.1061/(ASCE) Pugh, J. S.; Lowes, L.; and Lehman, D., 2015, “Nonlinear Line-Element
ST.1943-541X.0001083 Modeling of Flexural Reinforced Concrete Walls,” Engineering Structures,
Kolozvari, K.; Orakcal, K.; and Wallace, J. W., 2018, “New OpenSees V. 104, pp. 174-192. doi: 10.1016/j.engstruct.2015.08.037
Models for Simulating Nonlinear Flexural and Coupled Shear-Flexural Rejec, K., 2011, “Inelastic Shear Behaviour of RC Structural Walls under
Behavior of RC Walls and Columns,” Computers & Structures, V. 196, Seismic Conditions,” PhD dissertation, University of Ljubljana, Ljubljana,
pp. 246-262. doi: 10.1016/j.compstruc.2017.10.010 Slovenia, 347 pp.
Kolozvari, K., and Wallace, J. W., 2016, “Practical Nonlinear Modeling Saatcioglu, M., and Razvi, S. R., 1992, “Strength and Ductility of
of Reinforced Concrete Structural Walls,” Journal of Structural Engi- Confined Concrete,” Journal of Structural Engineering, ASCE, V. 118,
neering, ASCE, V. 142, No. 12, p. G4016001 doi: 10.1061/(ASCE) No. 6, pp. 1590-1607. doi: 10.1061/(ASCE)0733-9445(1992)118:6(1590)
ST.1943-541X.0001492 Scott, B. D.; Park, R.; and Priestley, M. J. N., 1982, “Stress-Strain
Los Angeles Tall Buildings Structural Design Council, 2017, “An Behavior of Concrete Confined by Overlapping Hoops at Low and High
Alternative Procedure for Seismic Analysis and Design of Tall Buildings Strain Rates,” ACI Journal Proceedings, V. 79, No. 1, Jan.-Feb., pp. 13-27.
Located in the Los Angeles Region,” Los Angeles, CA. Scott, H. M.; Franchin, P.; Fenves, G. L.; and Filippou, F. C., 2004,
“LS-Dyna,” 2018, Livermore Software Technology Group, Livermore, CA. “Response Sensitivity for Nonlinear Beam-Column Elements,” Journal of
Lu, Y., and Panagiotou, M., 2014, “Three-Dimensional Cyclic Beam- Structural Engineering, ASCE, V. 130, No. 9, pp. 1281-1288. doi: 10.1061/
Truss Model for Nonplanar Reinforced Concrete Walls,” Journal of (ASCE)0733-9445(2004)130:9(1281)
Structural Engineering, V. 140, No. 3, p. 04013071 doi: 10.1061/(ASCE) Spacone, E.; Filippou, F. C.; and Taucer, F. F., 1996, “Fiber Beam-Column
ST.1943-541X.0000852 Model for Non-Linear Analysis of R/C Frame: Part I. Formulation,” Earth-
Lu, Y.; Panagiotou, M.; and Koutromanos, I., 2014, “Three-Dimensional quake Engineering & Structural Dynamics, V. 25, No. 7, pp. 711-725. doi:
Beam-Truss Model for Reinforced-Concrete Walls and Slabs Subjected to 10.1002/(SICI)1096-9845(199607)25:7<711::AID-EQE576>3.0.CO;2-9
Cyclic Static or Dynamic Loading,” PEER Report 2014/18, Pacific Earth-

1656 ACI Structural Journal/November 2018


Stevens, N. J.; Uzumeri, S. M.; Collins, M. P.; and Will, T. G., 1991, Engineering & Structural Dynamics, V. 45, No. 13, pp. 2063-2083. doi:
“Constitutive Model for Reinforced Concrete Finite Element Analysis,” 10.1002/eqe.2731
ACI Structural Journal, V. 88, No. 1, Jan.-Feb., pp. 49-59. Vecchio, F. J., and Collins, M. P., 1986, “The Modified Compression
Taucer, F. F.; Spacone, E.; and Filippou, F. C., 1991, “A Fiber Beam- Field Theory for Reinforced Concrete Elements Subjected to Shear,” ACI
Column Element for Seismic Response Analysis of Reinforced Concrete Journal Proceedings, V. 83, No. 2, Mar.-Apr., pp. 219-231.
Structures,” Report No. UCB/EERC-91/17, Earthquake Engineering Vecchio, F. J., and Collins, M. P., 1993, “Compression Response of Cracked
Research Center, College of Engineering, University of California Berkeley, Reinforced Concrete,” Journal of Structural Engineering, ASCE, V. 119,
Berkeley, CA, 141 pp. No. 12, pp. 3590-3610. doi: 10.1061/(ASCE)0733-9445(1993)119:12(3590)
Thomsen, J. H., and Wallace, J. W., 1995, “Displacement-Based Design Vecchio, F. J., and Lai, D., 2004, “Crack Shear-Slip in Reinforced
of Reinforced Concrete Structural Walls: An Experimental Investigation Concrete Elements,” Journal of Advanced Concrete Technology, V. 2,
of Walls with Rectangular and T-Shaped Cross-Sections,” Report No. No. 3, pp. 289-300. doi: 10.3151/jact.2.289
CU/CEE-95/06, Department of Civil Engineering, Clarkson University, VecTor Analysis Software, 2013, Department of Civil Engineering,
Potsdam, NY, 344 pp. University of Toronto, Toronto, ON, Canada.
Tran, T. A., and Wallace, J. W., 2015, “Cyclic Testing of Moderate-As- Vintzeleou, E. N., and Tassios, T. P., 1987, “Behavior of Dowels under Cyclic
pect-Ratio Reinforced Concrete Structural Walls,” ACI Structural Journal, Deformations,” ACI Structural Journal, V. 84, No. 1, Jan.-Feb., pp. 18-30.
V. 112, No. 6, Nov.-Dec., pp. 653-665. doi: 10.14359/51687907 Vulcano, A.; Bertero, V. V.; and Caloti, V., 1989, “Analytical Modeling
Tsai, W. T., 1988, “Uniaxial Compressional Stress-Strain Relation of R/C Structural Walls,” Proceedings of the 9th WCEE, Tokyo-Kyoto,
of Concrete,” Journal of Structural Engineering, ASCE, V. 114, No. 9, Japan, pp. 41-46.
pp. 2133-2136. doi: 10.1061/(ASCE)0733-9445(1988)114:9(2133) Wallace, J. W., 2012, “Behavior, Design, and Modeling of Structural
Ulugtekin, D., 2010, “Analytical Modeling of Reinforced Concrete Walls and Coupling Beams - Lessons from Recent Laboratory Tests and
Panel Elements under Reversed Cyclic Loadings,” MS thesis, Department Earthquakes,” International Journal of Concrete Structures and Materials,
of Civil Eng., Boğaziçi University, Istanbul, Turkey, 140 pp. V. 6, No. 1, pp. 3-18. doi: 10.1007/s40069-012-0001-4
Vallenas, J. M.; Bertero, V. V.; and Popov, E. P., 1979, “Hysteretic Wallace, J. W.; Elwood, K. J.; and Massone, L. M., 2008, “Investigation
Behavior of Reinforced Concrete Structural Walls,” Report No. UCB/ of the Axial Load Capacity for Lightly Reinforced Wall Piers,” Journal of
EERC-79/20, Earthquake Engineering Research Center, University of Cali- Structural Engineering, ASCE, V. 134, No. 9, pp. 1548-1557. doi: 10.1061/
fornia, Berkeley, Berkeley, CA, 268 pp. (ASCE)0733-9445(2008)134:9(1548)
Vásquez, J. A.; De la Llera, J. C.; and Hube, M. A., 2016, “A Regularized Yassin, M. H., 1994, “Nonlinear Analysis of Prestressed Concrete Struc-
Fiber Element Model for Reinforced Concrete Shear Walls,” Earthquake tures under Monotonic and Cyclic Loads,” PhD dissertation, University of
California, Berkeley, Berkeley, CA, 195 pp.

ACI Structural Journal/November 2018 1657


Researchers, Maximize
Your ACI Membership

 Technical Presentations and Documents: Access to a vast


abstract library, online presentations, webinars, and
educational documents are often free for members

Attend Conventions: Provides opportunity to develop codes


and standards, learn about the latest technology, and network

 Call for Papers: ACI is accepting the submission of papers for


conventions, committees, chapters, and subsidiaries

ACI membership gives researchers and ACI Convention Schedule


educators unique opportunities to expand City Location Dates
their careers and further their education.
Québec Québec City
ACI events such as ACI’s biannual City, QC, Convention Centre
March 24-28,
conventions give participants the chance 2019
Canada & Hilton Hotel
to participate in the development of
Duke Energy
industry codes and standards, learn about Cincinnati, Convention Center October 20-24,
the latest in concrete technology, network OH, USA & Hyatt Regency 2019
with other concrete professionals, and Cincinnati
fulfill potential continuing education
Chicago/
requirements. ACI membership also Rosemont,
Hyatt Regency March 29-
allows professionals access to a vast O’Hare April 2, 2020
IL, USA
abstract library, online educational
Raleigh Convention
presentations, webinars, and ACI Raleigh, October 25-29,
Center & Raleigh
education documents at either no or NC, USA
Marriott
2020
minimal cost. ACI and its subsidiaries
are always accepting the submission
of papers for potential publication and
distribution. More information about
ACI’s conventions, technical presentations
and documents, and calls for papers can
be found at www.concrete.org.

For more information visit Concrete.org   


1 Appendix A

2 Table A.1 Comparison of simulated and measured response quantities based on experimental and

3 analytical load-deformation envelopes


Point on the Envelope Stiffness
@ Yielding @ Vmax @ Dmax Initial Secant at Yield

Dtot,Y VY Dtot,U VU VX Kini,tot Kini,flex Kini,sh KY,tot KY,flex KY,sh


(inch) (kips) (inch) (kips) (kips) (kips/in) (kips/in) (kips/in) (kips/in) (kips/in) (kips/in)
Test 0.80 30 3.15 36.5 36.5 91 726 24200 38 418 1874
RW2

Model 0.60 30 3.11 35.4 35.4 78 1000 957 50 600 957


Test 0.45 173 1.08 189.0 180.5 548 693 2766 386 693 874
SP4

Model 0.69 204 2.16 248.3 248.3 306 1206 410 298 1095 410
SW‐P3D

Test 0.90 44 5.00 50.0 32.0 88 ‐ ‐ 49 ‐ ‐


R2

Model 0.65 40 6.00 53.0 53.0 65 ‐ ‐ 62 ‐ ‐


WSH

Test 0.73 124 2.01 134.2 125.0 492 ‐ ‐ 170 ‐ ‐


Model 0.83 121 2.42 128.5 128.5 259 ‐ ‐ 147 ‐ ‐
Test 0.54 177 1.45 190.0 185.0 545 ‐ ‐ 334 ‐ ‐
S6

Model 0.54 170 1.97 218.0 218.0 340 ‐ ‐ 315 ‐ ‐


Test 0.80 30 3.15 36.5 36.5 91 726 24200 38 418 1874
SP4 RW2

Model 0.57 31 3.15 33.0 33.0 122 1830 7625 54 610 7625
Test 0.45 173 1.08 189.0 180.5 548 693 2766 386 693 874
Model 0.35 165 1.40 204.5 199.0 987 1410 7050 470 1175 914
Test 0.90 44 5.00 50.0 32.0 88 ‐ ‐ 49 ‐ ‐
FWE
R2

Model 0.90 42 6.00 53.0 53.4 139 ‐ ‐ 46 ‐ ‐


WSH

Test 0.73 124 2.01 134.2 125.0 492 ‐ ‐ 170 ‐ ‐


Model 0.58 110 1.90 125.0 85.0 660 ‐ ‐ 190 ‐ ‐
Test 0.54 177 1.45 190.0 185.0 545 ‐ ‐ 334 ‐ ‐
S6

Model 0.70 181 1.50 206.5 185.0 464 ‐ ‐ 259 ‐ ‐


Test 0.80 30 3.15 36.5 36.5 91 726 24200 38 418 1874
SP4 RW2

Model 0.78 29 3.15 35.3 35.3 74 618 3620 37 274 3621


Test 0.45 173 1.08 189.0 180.5 548 693 2766 386 693 874
SFI‐MVLEM‐FD

Model 0.50 186 1.09 186.0 167.5 1120 2135 2545 372 853 665
Test 0.90 44 5.00 50.0 32.0 88 ‐ ‐ 49 ‐ ‐
R2

Model 0.90 44 5.00 50.0 33.1 73 ‐ ‐ 48 ‐ ‐


WSH

Test 0.73 124 2.01 134.2 125.0 492 ‐ ‐ 170 ‐ ‐


Model 0.75 120 2.01 130.4 123.8 336 ‐ ‐ 161 ‐ ‐
Test 0.54 177 1.45 190.0 185.0 545 ‐ ‐ 334 ‐ ‐
S6

Model 0.54 174 1.50 187.0 185.5 731 ‐ ‐ 324 ‐ ‐


Test 0.80 30 3.15 36.5 36.5 91 726 24200 38 418 1874
SP4 RW2

Model 0.56 28 3.15 35.3 35.3 183 1650 16500 50 393 2357
Test 0.45 173 1.08 189.0 180.5 548 693 2766 386 693 874
SFI‐MVLEM‐SS

Model 0.29 165 1.44 204.5 193.0 1234 1410 9870 568 998 1337
Test 0.90 44 5.00 50.0 32.0 88 ‐ ‐ 49 ‐ ‐
R2

Model 0.90 42 6.00 53.0 53.0 199 ‐ ‐ 46 ‐ ‐


WSH

Test 0.73 124 2.01 134.2 125.0 492 ‐ ‐ 170 ‐ ‐


Model 0.78 117 3.70 136.5 136.5 574 ‐ ‐ 151 ‐ ‐
Test 0.54 177 1.45 190.0 185.0 545 ‐ ‐ 334 ‐ ‐
S6

Model 0.42 173 1.42 190.0 185.0 846 ‐ ‐ 420 ‐ ‐


Test 0.80 30 3.15 36.5 36.5 91 726 24200 38 418 1874
SP4 RW2

Model 0.58 29 3.19 33.0 33.0 166 1450 4350 51 483 1088
Test 0.45 173 1.08 189.0 180.5 548 693 2766 386 693 874
Model 0.33 165 1.44 184.9 157.5 1097 1974 2468 509 998 1045
NLTM

Test 0.90 44 5.00 50.0 32.0 88 ‐ ‐ 49 ‐ ‐


R2

Model 0.63 40 6.12 46.5 46.5 176 ‐ ‐ 64 ‐ ‐


WSH

Test 0.73 124 2.01 134.2 125.0 492 ‐ ‐ 170 ‐ ‐


Model 1.20 125 3.62 130.3 130.3 580 ‐ ‐ 104 ‐ ‐
Test 0.54 177 1.45 190.0 185.0 545 ‐ ‐ 334 ‐ ‐
S6

4 Model 0.60 206 1.00 206.0 182.5 495 ‐ ‐ 343 ‐ ‐

View publication stats

You might also like