Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

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

net/publication/351995204

Finite element analysis of rectangular and flanged steel‐concrete shear walls


under cyclic loading

Article  in  The Structural Design of Tall and Special Buildings · May 2021


DOI: 10.1002/tal.1863

CITATION READS

1 77

4 authors, including:

Ali Gharavi Siamak Epackachi


Sharif University of Technology University at Buffalo, The State University of New York
5 PUBLICATIONS   13 CITATIONS    64 PUBLICATIONS   470 CITATIONS   

SEE PROFILE SEE PROFILE

Rasoul Mirghaderi
University of Tehran
187 PUBLICATIONS   1,144 CITATIONS   

SEE PROFILE

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

Macro-Modeling of Reinforced Concrete Structures View project

Simulation of steel-concrete shear walls in high-rise building using PERFORM-3D View project

All content following this page was uploaded by Rasoul Mirghaderi on 18 April 2022.

The user has requested enhancement of the downloaded file.


Received: 8 April 2021 Accepted: 3 May 2021
DOI: 10.1002/tal.1863

RESEARCH ARTICLE

Finite element analysis of rectangular and flanged


steel-concrete shear walls under cyclic loading

Masoumeh Asgarpoor1 | Ali Gharavi1 | Siamak Epackachi2 |


Seyed Rasoul Mirghaderi3

1
Civil Engineering department, Sharif
University of Technology, Tehran, Iran Summary
2
Civil Engineering department, Amirkabir The seismic performance of steel-plate concrete (SC) shear walls is numerically inves-
University of Technology, Tehran, Iran
tigated for various SC wall geometrical shapes. Steel faceplates and infill concrete are
3
Civil Engineering department, University of
Tehran, Tehran, Iran two main components of SC walls; a full connection between steel and concrete is
required for composite action. A comprehensive set of SC wall specimens with differ-
Correspondence
Siamak Epackachi, Civil Engineering ent cross-sectional shapes and geometrical properties were tested in the literature
department, Amirkabir University of for evaluating their seismic behavior. The numerical responses of 17 tested SC walls,
Technology, Tehran, Iran.
Email: epackachis@aut.ac.ir including rectangular, T-shaped, and flanged walls, isolated and coupled walls, with
and without boundary elements and endplates, subjected to cyclic lateral loading and
a wide range of axial load ratio from 0.065 to 0.4, are evaluated using a finite
element-based software, LS-DYNA. The predicted force-displacement diagrams, fail-
ure modes, and damping ratios of the finite element models reasonably match the
test observations. Also, three main analysis and design issues are evaluated, including
the effect of coefficient of friction between concrete and steel faceplates, the effect
of modeling pullout of shear studs from the concrete on the global response of SC
walls, and the importance of the presence of shear studs and tie rods and the
possible changes that may happen by changing the connector spacing to faceplate
thickness ratio.

KEYWORDS
cyclic lateral loading, hysteresis loops, LS-DYNA, nonlinear finite element analyses, steel-plate
concrete (SC) shear walls

1 | I N T RO DU CT I O N A N D B A CK GR O U N D

Steel-plate concrete (SC) shear walls consist of an infill plain concrete covered by two steel faceplates on both sides. Tie rods, connecting bolts, or
batten plates are used as connectors between two steel plates, and headed studs or stiffening ribs attach the steel plates to infill concrete. These
walls are progressively being used in high-rise buildings for their less required cross-section area and better seismic performance than
corresponding conventional reinforced concrete or steel plate shear walls. Not only SC walls provide much energy dissipation through stable hys-
teresis cycles in seismic loading, but also they can be constructed more rapidly due to the elimination of rebars in concrete and the use of face-
plates as the permanent concrete framework.
In this paper, eight series of tested SC walls in the literature, containing a total number of 17 specimens, discriminated by their sectional
geometry, height to length ratio, and the existence of axial load, are selected for numerical investigation. Different geometrical shapes of walls
include rectangular, T-shaped, and flanged shapes, isolated and coupled, with and without endplates, with and without boundary elements. The
shape of the boundary elements can be circular or rectangular, constructed as HSS or build-up sections. The measured and DYNA-predicted

Struct Design Tall Spec Build. 2021;e1863. wileyonlinelibrary.com/journal/tal © 2021 John Wiley & Sons Ltd. 1 of 22
https://doi.org/10.1002/tal.1863
2 of 22 ASGARPOOR ET AL.

force-displacement responses are compared in terms of initial stiffness, peak strength, pinching in the cyclic loops, and strength degradation. The
failure modes of SC walls, including buckling and failure of faceplates, concrete cracking, and crushing, are evaluated at the end of tests. Energy
dissipation of wall models through hysteresis loops and equivalent viscous damping ratio are also derived and compared with those of test results.
Given the lack of robust concrete material in most FE programs, the fundamental characteristics of concrete material such as opening and
close of cracks, tension stiffening, strain softening, and shear-retention ability cannot be captured accurately. One of the main objectives of this
paper is the development of a reliable model to simulate the cyclic response of SC walls with a high level of accuracy in stiffness, strength, and
capturing pinching. For this purpose, a versatile finite element software LS-DYNA is used, and robust steel and concrete material models are pres-
ented to capture the actual behavior of SC walls. The robustness of modeling methodology is assessed by modeling 17 SC walls, including a vari-
ety of geometrical shapes, aspect ratio, and loading. Also, the effect of the friction coefficient between steel faceplates and concrete on the global
response of SC walls is investigated. Further attempts include the investigation of the effect of bonding between headed studs and concrete on
the global response of SC walls and the examination of the effect of the connectors arrangement.
Experimental studies were first conducted on concrete-filled steel structures (SC structures) to evaluate their resistance against seismic forces.
Sasaki et al.[1] performed bending shear tests on H-section SC walls with test variables, including shear span ratio, axial force, and reinforcement
ratio. They also compared the SC wall test results with those of reinforced concrete (RC) wall structures having the same reinforcement ratio, con-
cluding that SC walls showed greater yield strength and stiffness than those of RC walls. Usami et al.[2] surveyed the buckling characteristics of
the steel plates through experimental studies on SC elements under high compression loading. Takeda et al.[3] tested seven SC wall panels under
in-plane cyclic shear loading and investigated the effect of faceplate thickness, the number of partitioning webs, and the configuration of stud
bolts. These test series were subsequently continued by Ozaki et al.,[4] who also verified test results with analytical calculations and derived an
equation based on truss analogy to estimate the strength of SC panels. Suzuki et al.[5] used Sasaki's test data to develop a macro model for flanged
walls in order to reasonably calculate the maximum shear strength of SC walls. Vecchio and McQuade[6] have adopted the distributed stress field
model (DSFM), a smeared rotating crack model, to be used for the evaluation of SC wall structures. They have verified their theory using panel ele-
ments subjected to various loading conditions, including uniaxial compression, in-plane shear, and cyclic lateral loading. Although their studies
resulted in an accurate prediction of load capacity, post-peak ductility, and ultimate failure mode, they have not investigated a variety of SC wall
geometrical shapes. Moreover, their study has not been implemented to a commercial finite element software in order to be usable for other
researchers. The cyclic behavior of SC walls with steel sheets and steel-plate HSC were studied through experimental program.[7,8] Hu et al.[9]
experimentally investigated the seismic response of steel-plate composite frames strengthened with beam-only-connected SC walls. The results
of these tests showed higher stiffness and strength of frames and composite action between the composite frame and SC wall. Yan et al.[10] pro-
posed C-shaped connectors to attach the infill concrete to steel faceplates and experimentally investigated their axial compression strength. Guo
et al.[11] experimentally investigated the effect of out-of-plane eccentric axial compression, which induced out-of-plane bending moment, on the
in-plane cyclic response of SC walls and concluded that it significantly reduces the bearing and deformation capacity of specimens. Liew et al.[12]
suggested using lightweight steel-concrete composite structures as efficient modular systems that can be installed fast in high-rise buildings.
There are several available numerical studies in the literature to simulate the seismic behavior of SC shear walls. Ali et al.[13] investigated the
nonlinear behavior of four I-shaped SC walls through three-dimensional finite element modeling in ABAQUS. The pre-peak force-displacement
results showed a good agreement between numerical and experimental responses, although their post-peak behavior in monotonic loading and
finite element results of walls under cyclic loading were not addressed. Three sets of rectangular SC walls were numerically validated by Rafiei
et al.[14] under in-plane monotonic shear loading using ABAQUS along with a parametric study to investigate the effects of yield strength of the
steel plate, the compressive strength of the infill concrete, and the number of intermediate fasteners on the response of SC walls. Kurt et al.[15]
studied the response of rectangular and flanged SC walls subjected to monotonic lateral loading in LS-DYNA using finite element modeling and
presented recommendations for SC wall to basement anchorage design. Epackachi et al.[16] have developed finite element models in LS-DYNA for
predicting the cyclic response of four rectangular squat SC walls. Polat and Bruneau[17] conducted a numerical study on the previously tested rect-
angular concrete-filled sandwich steel panel walls (CFSSP) with half-round and round HSS boundary elements. They compared numerically
predicted and experimentally measured cyclic force-displacement diagrams, including initial stiffness, peak wall strength, and pinching in the hys-
teresis loops. Dey et al.[18] numerically investigated the seismic behavior of 4-storey and 6-storey composite shear walls using finite element
modeling in ABAQUS. They proposed equations to estimate the fundamental period of composite structures and shear stud spacing.[18] The
macro model of SC walls has been developed by Haghi et al.[19] using PERFORM-3D. Li et al.[20] analytically studied some concrete-filled double
steel plate shear walls using ABAQUS. An analytical study has been conducted to calculate the shear resistance of composite walls and columns
with encased steel profiles in concrete.[21] Haghi et al.[22] numerically investigated the effect of using boundary elements on stiffness, strength,
and energy dissipation of SC walls using finite element modeling in LS-DYNA. A variety of nonlinear finite element analyses in LS-DYNA were
conducted by Epackachi et al.[23] to propose a trilinear backbone curve and an equation for the peak strength of SC walls. Polat[24] studied the
percentage of steel plate yielding at different SC wall elevations, the distribution of cracks in concrete, and the contribution of SC wall compo-
nents in total base shear by modeling two tested SC walls in LS-SDYNA. Polat and Bruneau[17] provided some important factors for the design of
SC walls, including distribution of shear force, effective plastic strain at failure, and shear force demand in critical tie bars by finite element model-
ing in LS-DYNA. In the literature, there is a lack of a reliable finite element modeling procedure that guarantees a reasonable response for a wide
ASGARPOOR ET AL. 3 of 22

variety of SC walls. In this paper, a reliable finite element model is proposed to simulate the cyclic response of a wide range of previously tested
SC walls with different aspect ratios, section geometries, and axial force values. General recommendations for modeling different SC wall types,
including concrete and steel material models, element formulation, and contact definition between concrete and faceplates, are also presented.
Numerical models including a plastic-damage model for concrete, ignoring opening and closure of cracks and shear strength and stiffness
reduction over crack surface, cannot capture the pinching response of SC walls accurately. The developed model includes its simple and easy-
to-understand concrete and steel material models with minimum required input parameters. It can reasonably capture the fundamental
characteristics of SC wall behavior, such as cyclic behavior, pinching, energy dissipation, and failure modes for a variety of SC walls with different
design variables (e.g., cross-sectional shapes, aspect ratio, axial load, slenderness ratio, and reinforcement ratio).

2 | E X P E R I M E N T A L P R O G R A M FO R N U M E R I C A L V A L I D A T I O N

Eight groups of SC wall tests, each conducted by specific researchers, are chosen as indicators of all types of SC wall experiments for validating
the numerical models. The shape and geometric details of the total 17 SC wall specimens are summarized in Table 1. Epackachi et al.[25] tested
four large-scale rectangular SC wall specimens with an aspect ratio (height-to length, H/L) of 1.0 under displacement-controlled cyclic lateral load-
ing. The design variables include wall thickness, reinforcement ratio, stud, and tie bar spacing. The faceplates were welded to a 25.4-mm-thick
base plate bolted to a foundation block through 22 dywidag bars. Alzeni and Bruneau[26] tested four large-scale SC wall specimens with an aspect
ratio of 2.7 in two groups of two specimens, one group with boundary elements and another without boundary elements. Half circular endplates
and circular boundary elements were used instead of rectangular cap plates in order to avoid stress concentration at section corners. The differ-
ence between the specimens of each group is the spacing of the tie rods. The steel faceplates were continued down to the foundation and
secured within the foundation using web shear connectors. The base of the walls can be considered rigid in the numerical analysis due to the con-
tinuity of the faceplates into the foundation. The specimens Wall-2 and Wall-3 are chosen for numerical study in this paper as indicators of ele-
ments without and with boundary elements, respectively. Nie et al.[27] tested 12 SC wall specimens with square boundary elements. Two steel
faceplates were filled with high-strength concrete. The web plates were divided by continuous stiffeners, to which distributed batten plates were
transversely attached. Shear studs were also used for composite action between steel and concrete. The steel faceplates in all specimens
were welded to the base plate attached to a strong steel profile. All specimens were subjected to both high axial force and cyclic lateral loading.
The main design variables were thicknesses of the steel web plate and boundary elements, wall dimension and aspect ratio, the compressive
strength of concrete, and reinforcement ratio. Two wall cross-section dimensions were used in total; one of each is chosen for numerical study in
this paper. Cho et al.[28] tested six identical SC walls with rectangular cross-sections without endplates under cyclic lateral displacement. Out of
six specimens, three specimens were subjected to in-plane loading, and the other specimens were subjected to out-of-plane loading. One of the
specimens subjected to in-plane loading is selected for this study. Shafaei et al.[29] tested five rectangular walls with endplates under both axial
compressive force and lateral cyclic displacement. All specimens have the same geometric dimensions, and the design variables were tie rod spac-
ing, presence of studs, the ratio of axial compressive load, and concrete compressive strength. Two of the specimens with different values of the
design variables are used in this study.
Ge et al.[30] tested four rectangular walls, two with stiffening ribs and two with steel diaphragms under reversed lateral loading. The stiffening
ribs attach the faceplates to the infill concrete, and the steel diaphragms attach the faceplates together and divide the wall into two parts. Two
end plates were used only in the walls with wall-mid diaphragms, while both ends of the walls with stiffening ribs are open. The walls were
extended to the foundation, and the faceplates were connected to the rigid bottom plate through threaded bars. Lou et al.[31] tested nine rectan-
gular wall specimens under combined cyclic lateral and compressive axial loadings. The stiffener plates and through-thickness threaded rods were
used for connecting the faceplates. The design variables include wall height, the thickness of faceplates and stiffener plates, the presence of con-
necting bolts and shear studs, axial load ratio, and concrete compressive strength. Two of these walls with different aspect ratios and compressive
axial load ratios are selected for numerical study in this paper. Eom et al.[32] tested two rectangular and one T-shaped isolated walls together with
one rectangular and one flanged coupled walls under cyclic loading. Out of six tested walls, five specimens (Walls 13 to 17) are used for numerical
study in this paper, all of them were strengthened at their base with cover plates, except for the Wall-13. The brittle failure of the welded joint at
the wall base for specimen Wall-13 caused an early failure in that specimen. The presence of cover plates at wall bases transmits the location of
local buckling of steel faceplates to a distance far from the base. In Wall-17, the connections of the coupled beams with the wall plates were also
strengthened by cover plates to avoid the weld failure of connections, as occurred in Wall-16.

3 | N U M E R I C A L M O DE L I N G O F S C WA L L S

All numerical analyses are conducted using the versatile finite element software LS-DYNA.[33] The modeling assumptions, including material
models, boundary conditions, and element types, are discussed as follows.
TABLE 1 Test specimen configurations
4 of 22

Stud
Specimen Wall Wall dimension (H  L  T) spacing Tie rod/batten plate Faceplate Axial compressive Description of the cross-sectional
Researcher ID type (mm  mm  mm) (mm) spacing (mm) thickness (mm) load (kN) shape (mm)
Epackachi Wall-1 Isolated 1524  1524  305 102 305 47 - Rectangular
et al.[25] Wall
Alzeni and Wall-2 3048  1235  219 - 203 7.94 - Rectangular with half-round HSS
Bruneau[26] 219  8.18 end plates
Wall-3 3048  1059  168 - 203 7.94 - Rectangular with round HSS
219  8.18 boundary elements
Nie et al.[27] Wall-4 2568  1284  214 100 145 4 7535 Rectangular with Box 214  214  4
boundary elements
Wall-5 1125  750  125 100 85 3 2718 Rectangular with Box 125  125  4
boundary elements
Cho et al.[28] Wall-6 1100  400  150 200 - 3 - Rectangular
Shafaei Wall-7 2743  914  279 - 114 4.76 934.5 Rectangular with PL. 279  4.76 end
et al.[29] plates
Wall-8 2743  914  279 114 228 4.76 2403 Rectangular with PL. 279  4.76 end
plates
Ge et al.[30] Wall-9 1420  800  225 75 150 3.75 - Rectangular
Wall-10 1420  800  225 75 3.75 - Rectangular with PL. 225  3.75 end
plates
Luo et al.[31] Wall-11 2160  1200  120 160 160 8 1601 Rectangular with PL. 120  8 end
plates
Wall-12 3000  1200  120 160 160 6 1040 Rectangular with PL. 120  6 end
plates
Eom et al.[32] Wall-13 3850  1000  120 - 150 10 - Rectangular with PL. 120  10 end
plates
Wall-14 3850  1000  120 - 150 10 - Rectangular with PL. 120  10 end
plates
Wall-15 3850  1000  120 - 150 10 - T-shaped
Wall-16 Coupled 4000  1500  100 - 150 5.9 - Rectangular with PL. 100  5.9 end
Wall plates
Wall-17 4000  1500  600 - 150 5.9 - Flanged
ASGARPOOR ET AL.
ASGARPOOR ET AL. 5 of 22

3.1 | Concrete material model

Among various developed concrete material models, the Winfrith model, based on the combination of the plasticity and smeared crack approach,
is used to model concrete. Previous studies have proven that the Winfrith model can reasonably simulate the cyclic behavior, peak strength, and
pinching of the reinforced concrete structures, in which the response mainly depends on the concrete behavior.[16,34] The plasticity model pro-
posed by Ottosen[35] considering the non-circular meridian and deviatoric section for concrete has been implemented in the Winfrith model. The
Winfrith model considers a linear tensile response up to the peak tensile strength and a linear or bilinear post-peak strength degradation if strain
rate effects are ignored or not, respectively. The strain rate effect is commonly ignored to simulate quasi-static experiments by setting the value
of the parameter “RATE” equal to 1. The area under the curve of post-peak tensile strength versus the crack width is known as fracture energy
and can be calculated using Equation 1 based on the CEB-FIP Model Code[36] in the absence of experimental data.

GF ¼ 73:f cm 0:18 ð1Þ

where GF is the fracture energy in N/m, and fcm is the mean value of the compressive strength in MPa. The crack width at which tensile stress nor-
mal to the crack goes to zero can be determined as 2Gf =f 0t , in which f 0t is maximum tensile strength, as shown in Figure 1. The crack width can be
implemented to the model through the parameter “FE.” The smeared crack approach developed by Broadhouse[38] enables the concrete to bear
force after cracking. The Winfrith model can show the crack patterns on deformed shape over the entire loading process. For concrete elements
under compression loading, the elements continue to resist the applied load until an erosion criterion prevents the contribution of that element in
load-bearing. Herein, the minimum principal strain (MPS) is considered as erosion criterion for concrete elements, implying that a concrete ele-
ment is crushed and thus is no longer contributes to load-bearing in the system when the MPS of that element reaches the considered threshold.
The concrete modeling parameters for all SC walls are presented in Table 2. In this table, for walls with no significant post-peak strength degrada-
tion or concrete crushing, the concrete failure strain has not been included in the analysis model. Based on the results of the numerical analyses
and calibration of the concrete failure strain, in the absence of test data and for preliminary analysis of SC walls, the value of the failure
(i.e., crushing) strain of concrete elements can be taken as ten times the crushing strain of unconfined normal concrete.

3.2 | Steel material model

An elastic–plastic material model is used for steel material modeling. The steel material model considers both isotropic and kinematic hardening
rules through a parameter β, as shown in Figure 2.[39] In this figure, the true tensile stress is illustrated versus true tensile strain. The parameter
E is Young's modulus, and Et is the tangent modulus (i.e., the slope of the plastic part of the stress–strain curve). The coupon test data are used to
calculate the yield stress Fy, and Et using the equivalent energy method. In this paper, the kinematic behavior is assumed for the steel material by
setting the β value equal to 0 in all analyses. This model considers a continuous increase in strength of elements unless the effective plastic strain
(i.e., the cumulative plastic strain over the entire load stages) reaches the considered threshold defined as parameter failure strain (FS) in

FIGURE 1 Tensile stress versus tensile displacement when strain rate effects are not considered[37]
6 of 22 ASGARPOOR ET AL.

TABLE 2 Material properties input to the DYNA models

Steel properties Concrete properties

Uniaxial
Young's Yield Tangent Effective Young's Uniaxial tensile Crack
Specimen modulus strength modulus plastic strain modulus compressive strength width Erosion
ID (MPa) (MPa) (MPa) at failure (MPa) strength (MPa) (MPa) (mm) strain
Wall-1 200,000 320.6 231 0.9 20,684 31 2.4 0.064 0.04
Wall-2 200,000 455 296 3.5 27,647 47.85 3.97 0.053 -
Wall-3 200,000 460.6 324 - 32,819 48.8 4.02 0.053 -
Wall-4 200,000 434 345 1.35 49,366 89.8 6.05 0.053 -
Wall-5 200,000 497.8 27,579 0.3 56,192 80.67 5.63 0.053 -
Wall-6 200,000 337.8 4371 0.15 27,785 35 3.23 0.052 0.035
Wall-7 200,000 447.5 689.5 1 31,509 44.87 3.8 0.053 -
Wall-8 200,000 447.5 257.2 0.6 35,783 57.98 4.5 0.048 -
Wall-9 200,000 255.1 6894 - 24,614 27.5 2.75 0.051 0.04
Wall-10 200,000 255.1 275.8 2.2 24,614 27.5 2.75 0.051 0.08
Wall-11 200,000 276.8 2758 0.23 30,605 32.5 3.07 0.052 -
Wall-12 200,000 351.6 5515.8 0.3 30,605 32.5 3.07 0.052 0.05
Wall-13 200,000 382.6 689.4 - 29,647 39.7 3.51 0.052 -
Wall-14 200,000 382.6 6894 0.36 25,200 39.88 3.51 0.052 -
Wall-15 200,000 382.6 6894 0.3 29,647 39.7 3.51 0.052 -
Wall-16 200,000 371.9 5515 0.28 35,990 58.69 4.55 0.053 0.05
Wall-17 200,000 371.9 6894 1 35,990 58.69 4.55 0.053 -

FIGURE 2 Elastic–plastic behavior of steel material model with kinematic and isotropic hardening rules[39]
ASGARPOOR ET AL. 7 of 22

LS-DYNA. Once the effective plastic strain of one steel element reaches the FS value, that element erodes, and no more contribute to the
strength of the SC wall. This parameter has an essential influence on the post-peak response of SC walls. This failure strain of steel faceplates
depends on wall geometry, material properties, and cyclic loading history. Walls with a larger number of load cycles have larger effective plastic
strain. This value is obtained by the calibration of numerical values with test results for each SC wall. In the absence of test provided information
for yield and ultimate strengths of tie rods and studs, the same material properties of faceplates are used without defining failure strain. The input
parameters for steel material are summarized in Table 2. In the cases that no post-peak response is recorded by the tests, no failure strain is con-
sidered for the steel material, as seen in Table 2. In addition, the values of the Poisson's ratio and maximum aggregate size are considered to be
0.18 and 19 mm, respectively.

3.3 | Element modeling

Eight noded solid elements with reduced integrated (RI) element formulation, which considers constant stress in mid-point of the elements,
are used for concrete and base plate modeling. The use of RI element formulation results in faster and more efficient analyses than fully
integrated element formulation, but it may cause zero energy deformation modes in solid elements. This issue can be solved using proper
hourglass type and coefficient for stabilization. In this paper, the stiffness form of Flanagan-Belytschko hourglass (type 5) with a coefficient
value equal to 0.05 is applied to all solid elements.[33] Four noded shell elements with Belytschko-Tsay element formulation, which can con-
sider local buckling of steel faceplates, are used for modeling them.[33] Tie rods and shear studs are modeled using beam elements with
Hughes-Liu element formulation.[33] In many of the test specimens, the base plate is attached to the strong floor through a number of post-
tensioning bars, which are modeled with translational elastic springs using discrete elements. The stiffness coefficient of the springs can be
calibrated in a way that the initial stiffness of the model matches the test value. The LS-DYNA models of all specimens are depicted in
Figure 3.

3.4 | Contact and constraint modeling

To avoid the penetration of the elements to each other, Contact_Automatic_Surface_To_Surface is defined in the models between the infill con-
crete and the faceplates and the infill concrete and baseplates.[33] This contact allows the transfer of loads between the master and slave seg-
ments by means of defining a friction coefficient. Based on several tests conducted by Rabbat and Russell,[40] it is recommended to take the
coefficient of static friction for concrete cast on steel plate a value between 0.57 and 0.7. Since the use of different values makes no significant
change in global results, a mean value of 0.5 is used for all specimens in this paper. The contact between the steel faceplates and baseplate is
defined by Contact_Tied_Shell_Edge_To_Surface, available in LS-DYNA. A full bond between the nodes of stud and tie bar, and the concrete
is assumed using Constrained_Lagrange_In_Solid in the cases that the nodes do not lie in the same coordinate, so it is not possible to merge the
nodes with the same coordinates. Defining the constraint ignores the slippage between beam elements and solid elements. The effect of bonding
between shear studs and concrete on the response of SC walls is also examined later.

4 | ANALYSIS RESULTS

4.1 | Force-displacement hysteresis response

The force-displacement responses of test and finite element models are compared with each other for all 17 specimens in Figure 4. The
values of the lateral force and displacement are recorded as the shear reaction of all base nodes and the lateral displacement at the mid-
height of the loading plates, respectively. The analysis results show a reasonable agreement with the test results for all types of tested SC
walls, including isolated and coupled walls, rectangular, T-shaped, and flanged walls, walls with or without endplates, boundary elements, and
axial compressive load. The initial stiffness, peak strength, post-peak strength degradation, unloading/reloading stiffness, and pinching compare
acceptably with the test results. The initial stiffness of those walls with baseplate wall-to-foundation connection could be captured with
modeling linear springs under the baseplate. The peak strength of specimens is sensitive to the values of the yield strength and tangent mod-
ulus of the steel material. Post-peak strength degradation of the test results can be captured by defining specific values for effective plastic
strain and minimum principal strain as erosion criteria for steel and concrete elements, respectively. The ability of the Winfrith concrete mate-
rial in considering the energy dissipation through opening and closing of cracks during loading and unloading, together with the ability of the
steel material model in capturing the buckling of steel faceplates, cause the numerical models to favorably simulate pinching as occurred in
experiments.
8 of 22 ASGARPOOR ET AL.

FIGURE 3 LS-DYNA model of SC walls


ASGARPOOR ET AL. 9 of 22

FIGURE 4 Test and numerical force-displacement responses


10 of 22 ASGARPOOR ET AL.

4.2 | Hysteresis damping ratio

The equivalent viscous damping ratio ξeq for each cycle of SC walls can be calculated using Equation 2.[41]

1 ED
ξeq ¼ ð2Þ
4π ES0

where ED is the dissipated energy of each enclosed loop, and ES0 is the strain energy, and is the area of a triangle beneath a tangent line con-
necting the origin and maximum drift points on the curve. The comparison between the hysteresis damping ratios of tests and models is made in
each drift ratio, which clarifies the accuracy of pinching simulations; see Figure 5. Although some of the models provide accurate results, there is
not a similar trend for all walls. The hysteresis damping ratio is underestimated or overestimated for some walls. The hysteresis damping ratio
is calculated using the area under the cyclic loops, so pinching has a significant effect on its value. The discrepancies between predicted and mea-
sured hysteretic damping is attributed to error for prediction of pinching in SC walls, which is expected as there are a number of behavioral factors
affecting pinching such as concrete cracking and crushing, and buckling and tearing of the steel faceplates, and small errors in predicting any of
those items lead to a considerable difference between measured and predicted pinching. However, based on the analysis results, the captured
pinching is still acceptable.

4.3 | Failure modes of SC walls

The DYNA-predicted failure modes of all 17 SC walls are compared with their corresponding test observations, as shown in Figure 6. In the range
of aspect ratio of SC walls considered in this study, the dominant behavior of all SC walls was flexure and plastic hinges formed at two end corners
of the wall base. The sequence of damage to most of the studied SC walls is concrete tensile cracking (A), buckling of the steel faceplates (B), con-
crete crushing (C), and fracture of the steel faceplates (D), as schematically presented in Figure 7.
The use of smeared cracking Winfrith model for concrete enables the display of the crack patterns of the infill concrete during the loading
process. The Winfrith model considers the strength and stiffness degradation of concrete elements after the first crack and redistributes the ele-
ment stresses until cracking happens in the other two principal directions. This approach realistically models the energy absorption of the concrete
elements during the opening and closing of the cracks. Steel faceplates buckle at the wall bases, where the effective plastic strains are concen-
trated, as illustrated in Figure 6. The shell elements reasonably capture the buckling of the steel faceplates as occurred in the experiments. The
concrete crushing occurs after steel buckling, especially in the walls with no end plates and boundary elements. In most of the SC walls, the lateral
strength reaches its maximum value when concrete crushing occurs. Concrete crushing is followed by steel fracture, and the post-peak strength
degradation of SC wall response is initiated. The faceplate shell elements are eroded when their effective plastic strain reaches the threshold
value, in which post-peak strength degradation begins. This threshold value depends on the wall geometry, material properties, and cyclic loading
history. As seen in Figure 6, the obtained results, including concrete cracking, steel buckling, concrete crushing, steel fracture, and the concentra-
tion of plastic strains, are predicted by SC wall models accurately.

5 | EVALUATION OF SOME MODELING AND DESIGN ISSUES

Herein, the effects of some modeling assumptions on the global response of SC walls are investigated. These major assumptions include the
effects of the shear friction between concrete and steel faceplates, bonding between the shear studs and concrete, and configurations of the tie
rods and studs over the web of the SC walls.

5.1 | Effect of friction between concrete and steel faceplates

Friction is considered as a force transfer mechanism between the concrete and steel faceplates before the sliding begins. Several studies investi-
gated the load transfer mechanism between steel and concrete via friction through various experimental programs as follows.
Some sets of push-off tests were conducted by Rabbat and Russell[40] on concrete or grout blocks with wet or dry interface over steel plates.
Based on their experiments, a range of static coefficients of friction (COF) from 0.57 to 0.7 was suggested under normal compressive stress. A
series of experiments was conducted by Evirgen and Tuncan[42] to investigate the shear strength of the interface layer between concrete and
steel in concrete-filled steel tubular columns. They concluded that the static friction coefficient between concrete and steel plates could be con-
sidered to be 0.55. The friction between the steel baseplate and the concrete was also explored through shaking table study for seismic
ASGARPOOR ET AL. 11 of 22

FIGURE 5 Test and numerical hysteresis damping ratios


12 of 22 ASGARPOOR ET AL.

FIGURE 6 Damage to the test specimens and contours of effective plastic microstrain in numerical models
ASGARPOOR ET AL. 13 of 22

FIGURE 7 Schematic description of failure sequence in SC walls

applications by McCormick et al.[43] They determined a constant friction coefficient equal to 0.78 for seismic load applications. Another series of
studies resulted in an average range of 0.52–0.97 for shear friction coefficient between steel baseplate and mortar in column bases.[44] To investi-
gate the effect of different suggested friction coefficient values between concrete and steel, five modeled walls with different design variables,
including Walls 1, 4, 10, 11, and 17, are reanalyzed with different contact friction values, including 0.3, 0.5, 0.7, and 1. Based on analysis results,
the friction between steel and concrete has a negligible effect on the global response of SC walls, implying that friction between the concrete
panel and steel faceplates in SC walls is not the effective parameter in the overall performance of SC walls. This result can be attributed to the
use of headed studs and threaded rods, preventing the sliding of steel faceplates over the concrete panel, so the friction is not activated.

5.2 | Effect of bonding between headed studs and concrete

In all SC wall models analyzed previously, it was assumed that the shear studs were fully bonded within the infill concrete using a
Constraint_lagrange_In_Solid keyword in the models. If shear studs slip inside the concrete or have inadequate embedment length for the genera-
tion of composite action between steel faceplates and infill concrete, the full constraint assumption may not work, and probably shear studs pull
out from the concrete, which may affect the global response of SC walls. To investigate whether and to what extent bonding between headed
studs and concrete may change the behavior of the SC walls, further analyses are performed using a nonlinear shear force-slip relationship
available in the literature for headed studs embedded in concrete. There are extensive experimental studies on the tensile and shear response of
headed studs embedded in concrete in the literature.[45–48] Based on available test data, Equation 3 was proposed as a shear force-slip relation-
ship for headed studs embedded in concrete (see Figure 8).[46] It should be noted that the proposed nonlinear force-displacement diagram is valid
for lightweight and normal-weight concretes. The details for the range of uniaxial compressive strength of concrete and anchor diameter are
provided in Ollgaard's study.[46]
 0:4
Q ¼ Qu 1  e18Δ ð3Þ

where Q is the shear load applied to the headed studs, Qu is the maximum shear strength of the embedded studs and can be calculated using
Equation 4, and Δ is the slip in inches. In this equation, the function yields 99% of the ultimate load for a slip equal to 0.2 in.
qffiffiffiffiffiffiffiffiffi
Qu ¼ 0:5As f 0c Ec ð4Þ

where As is the cross-sectional area of shear studs (in.2), f 0c is the uniaxial compressive strength (ksi), and Ec is the elastic modulus of concrete (ksi).
14 of 22 ASGARPOOR ET AL.

F I G U R E 8 Force-slip relationship of shear studs embedded in


normal concrete[46]

The above equations were validated by Zhang et al.[48] using another series of experiments.[45,47] They conducted finite element-based
numerical analyses in two different ways; first, using solid elements for modeling all components of specimens, including concrete, steel face-
plates, and shear studs attached to concrete with frictionless contact, and second, using the solid, shell, and beam elements for modeling concrete,
steel faceplates, and shear studs, respectively. In the second way, the force-slip relation of Equation 3 was applied to connector elements. They
concluded that both modeling types result in the same responses, and thus, Equation 3 can be implemented as the force-slip relation for beam
connector elements.
For this purpose, the MAT_Nonlinear_Plastic_Discrete_Beam (MAT_068) is defined for applying the force-slip curve of Figure 8 to the con-
nectors modeled by discrete beam elements. This definition enables the model to consider the pullout of shear studs from concrete during the
loading process. The effect of bonding between headed studs and concrete is investigated just in four walls, which have headed studs.
The obtained results show that modeling the pullout of studs from the concrete makes no changes in the force-displacement response of these
SC walls. It should be noted that the behavior of most of the SC walls are flexure-critical and plastic hinges form at the corners of the wall base.
Since the concrete crushing concentrates on a small region of the wall base, even if some of the studs pullout from the concrete, it cannot change
the overall response of these walls.

5.3 | Effect of the configuration of connectors over the web of SC walls

In some previous studies, the slenderness ratio, S/ts, has been considered as one of the major design variables,[25,26] where the parameter S is the
maximum spacing between two adjacent connectors and ts is the thickness of the steel faceplate. These tests concluded that the slenderness ratio
did not affect the peak resistance of their walls. The influence of this design variable on the maximum strength of SC walls has been studied by
the analytical studies as well.[23] The AISC-341[49] presents an equation for the maximum spacing of tie bars in the vertical and horizontal direc-
tion based on the results of Alzeni's study.[26] There is still a need for further investigation of the effect of using only studs or different configura-
tions of both studs and tie rods, which may not necessarily satisfy the limitations of the AISC-341 limitations. The available experiments in the
literature have used a conservative connector configuration and spacing.
In this paper, the effect of connector arrangement on SC wall response is considered by changing two variables: elimination of shear studs or
tie rods and changing the slenderness ratio S/ts. The results of studies conducted in this section are presented in Figure 9 and Table 3. The param-
eter λ in the legend of Figure 9 denotes the S/ts ratio of the numerically validated specimen, and any change to the ratio of S/ts is specified by a
coefficient multiplied to the parameter λ. The evaluation of Wall-1 shows that the removal of shear studs significantly reduces the strength of the
wall, while the removal of tie rods has a negligible effect on the response of this wall, as shown in Figure 9a. This result can be attributed to a con-
siderable number of shear studs used in Wall-1 in which their removal affects the composite action significantly, and if they remain, they will
cover the removal of tie rods. The use of small spacing between shear studs causes the concentration of the failure modes in the region between
the wall base and the first row of shear studs. In this wall, the results of multiplying the S/ts ratio by 0.5, 2, 3, and 4 are illustrated in Figure 9b.
ASGARPOOR ET AL. 15 of 22

F I G U R E 9 Force-displacement diagrams of SC walls with different


connector configurations
16 of 22 ASGARPOOR ET AL.

TABLE 3 Key results of SC walls with different connector configurations

Peak Crack Post-peak


Specimen strength Displacement corresponding Displacement at 20% strength stiffness
ID Model (kN) to peak strength (mm) strength loss (mm) (kN) (kN/mm)
Wall-1 Reference model 1568 13.8 24.4 470 25.2
Model without studs 938.7 21.3 40.7 282 11
Model without tie rods 1635 14.4 20.5 490 35.5
S/ts = 0.5λ 1625.3 13.4 22.3 487 34.7
S/ts = 2λ 1477 14.1 27 443 25.3
S/ts = 3λ 1344.7 14.1 32.2 403 13.5
S/ts = 4λ 836.5 14.3 43 250 1.2
Wall-2 Reference model 1354.5 53.8 89.2 406 7.2
S/ts = 2λ 1289 51.6 105.8 386 4.7
Model with replacing all 1350 54.2 89.6 405 6.9
tie rods with studs
Wall-4 Reference model 2208 32.8 46 662 28.9
Model without studs 2016 41.2 45.4 605 51.4
Wall-6 Reference model 128 56.6 67 38.4 2.5
Model with replacing all 134.2 55.9 70 40.3 2.2
tie rods with studs
S/ts = 0.5λ 147.5 42.4 55 44.2 2.8
Wall-8 Reference model 917.4 34.2 45.2 275.2 16.8
Model without studs 766.4 37.2 54 230 4.3
Model without studs 742.5 35.1 54 222 5.7
and tie rods
Wall-9 Reference model 560 32.2 60.1 168 2.5
Model without studs 466.7 32.6 58.5 140 0.3
Wall-10 Reference model 555.5 60.1 - - -
Model without studs 263.8 60.2 - - -
Wall-11 Reference model 1250.6 19.5 19.8 375.2 7
Model without studs 1121.8 19.4 20 336.5 2.7
Model without tie rods 1213.8 19.2 20.7 364 2.4
Model without studs 905.7 19.5 19.4 271.7 2.2
and tie rods
S/ts = 2λ 1246 19.5 20.5 373 3.3
Wall-14 Reference model 867.3 74.6 102 260 5.9
Model without tie rods 668 57.9 80.5 200 5.88
Model with replacing all 881.8 96.2 110 264 11.5
tie rods with studs
S/ts = 3λ 804.4 96.2 120 241 6.9

The peak strength of SC wall is not changed significantly when the ratio of S/ts is multiplied by 0.5 and 2. The strength reduction increases as S/ts
ratio is multiplied by larger factor of 3 and 4. The response of Wall-1 with S/ts = 4λ is similar to the response of this wall with no studs as shown
in Figure 9a.
In Wall-2, all tie rods are replaced by shear studs. As shown in Figure 9c, the peak strength is not affected, and a decrease in post-peak
strength can be observed. Doubling the spacing of the tie rods makes no significant change in the response of this wall, and since only tie rods are
used in this wall, no more increase in the S/ts ratio is possible. In Wall-4, the use of boundary elements and stiffeners with batten plates in the wall
length and height makes a strong connection between the concrete and steel faceplates and reduces the dependency of the results to the stud
spacing; the removal of all studs reduces the strength of this wall about 10% as shown in Figure 9d. In Wall-6, only shear studs are used as con-
nectors in the rectangular wall without endplates. The effect of shear stud replacement with tie rods in Wall-6 is shown in Figure 9e. Reducing
ASGARPOOR ET AL. 17 of 22

the S/ts ratio by 50% in this wall causes a 20% increase in peak strength. In Wall-8 with endplates, the removal of shear studs causes a 15% reduc-
tion in peak strength. These studs are the only connectors attaching concrete to the steel faceplates, but two endplates used in a relatively short
distance prevent the sudden buckling of the faceplates. The endplates are also attached to the concrete through a number of tie rods. As shown
in Figure 9f, the elimination of these tie rods, in addition to studs, do not affect the wall behavior significantly in comparison with the wall without
shear studs.
One important factor causing the composite action even after stud deletion is the presence of axial compressive load, which completely
enhances the confinement of concrete in the faceplates. Wall-9 is a rectangular SC wall without endplates, in which the concrete is attached to
the faceplates through stiffening ribs and shear studs. As shown in Figure 9g, the removal of studs reduces the peak strength by about 20%, which
again emphasizes the importance of the presence of connectors in open rectangular walls. In Wall-10, despite the presence of endplates and a
continuous plate as a diaphragm in the middle of the wall, the elimination of the studs significantly reduces the strength of the wall as shown in
Figure 9h because its thin faceplates with a thickness of 3.75 mm vulnerably buckle when their S/ts value increases. Also, there is no axial load to
confine the concrete between two steel faceplates. Both studs and tie rods are used in combination with the continuous stiffeners in the rectan-
gular Wall-11 with endplates, whose results are shown in Figure 9i. In this wall, doubling the S/ts value, like eliminating all tie rods, causes a small
change in wall behavior. By elimination of all studs and tie rods that can be achieved by multiplying the S/ts value by 4, about 27% reduction in
the peak strength of the wall happens. In Wall-14 with endplates, which has only tie rods as faceplate connectors, all tie rods are replaced by the
shear studs. This model shows the same strength as the original model but with less ductility and more post-peak strength reduction, as illustrated
in Figure 9j. The same behavior is also observed when the S/ts is multiplied by 3. When all tie rods are removed, the wall strengths in pre- and
post-peak regions of the wall are decreased significantly.
To conclude, there are several parameters that influence the effect of S/ts value on the global response of SC walls. Based on the analysis
results, the presence of axial compressive load on a concrete panel surrounded by any type of steel jackets such as endplates, stiffeners, and steel
diaphragms through the wall height of SC wall provides lateral confinement on concrete and hence its compressive strength and ductility are sig-
nificantly improved, which in turn the post-peak strength degradation and concrete failure are delayed.

6 | C O N C E P T U A L E V A L U A T I O N O F T H E BE H A V I O R O F S C W A L L S

6.1 | Peak strength

The peak lateral strength of any structural system is a key design parameter that indicates the capacity of the system to resist the applied lateral
loads. The plastic stress distribution (PSD) method is a common method for predicting the peak flexural strength of SC walls. AISC 341-16[49]
introduces this methodology for the calculation of the flexural strength of only two shapes of SC walls, including SC walls without and with
boundary elements. In this subsection, the validity of the PSD method for predicting the peak lateral strength of SC walls is investigated using the
results of 42 SC walls with four cross-sectional shapes tested by eight researchers in the literature. The wall types include double skin plates SC
walls (DS-SC walls), double skin plates with end plates SC walls (DSE-SC walls), SC wall with boundary elements (BE-SC walls), and flanged
SC walls (FL-SC walls). The ratio of the peak strength obtained by tests (VTest) to the peak strength obtained by AISC 341-16 formulation (VAISC)
using the PSD method is specified in Table 4.
Comprehensive numerical studies in this paper indicate that for numerical simulation and design of SC walls, the most important design vari-
ables include the wall shape, aspect ratio, steel yield strength, concrete compressive strength, slenderness ratio, reinforcement ratio (ratio of steel
area to total area in the section), and axial load ratio, which is defined in Equation 5.

P
N¼ ð5Þ
As F y þ Ac f 0c

where P is the axial load, As is the steel area, and Ac is the concrete area. The values of the peak strength ratio with consideration of design vari-
ables in Table 4 reveals notable conclusions as follows.
The AISC 341-16 formulation for calculation of the peak flexural strength is recommended to be used only for rectangular shapes as it calcu-
lates the peak strength of flanged walls with a considerable error. According to AISC 341-16 provision, the nominal moment strength of rectangu-
lar SC walls without boundary elements is equal to the yield moment, My, assuming a linear elastic stress distribution with maximum concrete
compressive strength limited to 0:7f 0c and maximum steel stress limited to Fy. This concept originates from the study of Kurt et al.,[50] which is
valid only for SC walls with an aspect ratio of less than 1.5. As seen in Table 4, the AISC 341-16 formulation predicts the shear strength of low
aspect ratio walls of Epackachi et al.[25] with good accuracy, but its peak shear strength predictions for double skin plate walls with large aspect
ratio, such as the walls tested by Cho et al.[28] and Ge et al.,[30] are not accurate. It is well known that the walls with a large aspect ratio are flexural
controlled, whereas, in low aspect ratio walls, shear behavior and the interaction of shear and flexure are dominant. It seems that the use of My
TABLE 4 Test specimen configurations

Specimen Aspect Steel yield strength Concrete compressive strength Slenderness Reinforcement Axial load ratio Peak strength
18 of 22

Researcher ID Shape ratio (MPa) (MPa) ratio ratio (%) (%) ratio
H Fy f 0c S As N V Test
L ts Ag V AISC

Epackachi et al.[25] SC-1 DS-SC 1 262 31 63.5 3.1 0 1.08


SC-2 walls 1 262 31 31.7 3.1 0 1.17
SC-3 1 262 36.5 47.5 4.2 0 1.05
SC-4 1 262 36.5 23.8 4.2 0 1.07
Cho et al.[28] SC-5 2.75 337.7 35 66.7 4 0 1.37
SC-6 2.75 337.7 32.5 66.7 4 0 1.31
SC-7 2.75 337.7 32.5 66.7 4 0 1.31
[30]
Ge et al. SC-8 1.8 255 27.5 54.5 2.4 0 1.43
SC-9 1.8 255 27.5 40 3.3 0 1.47
[30]
Ge et al. SC-10 DSE-SC 1.8 255 27.5 54.5 3.4 0 1.47
SC-11 walls 1.8 255 27.5 40 4.7 0 1.38
[29]
Shafaei et al. SC-12 2.95 422 44.8 23.7 5.2 6.56 1.22
SC-13 2.95 422 53.7 23.7 5.2 14 1.32
SC-14 2.95 422 50.9 23.7 5.2 20.4 1.33
SC-15 2.95 422 60.3 47.4 5.2 14.3 1.24
SC-16 2.95 422 58 47.4 5.2 14.2 1.24
Eom et al.[32] SC-17 3.85 383 39.7 30 18.3 0 1.06
SC-18 3.85 383 39.7 30 18.3 0 1.07
Alzeni and SC-19 2.5 434.4 47.9 25.4 8.25 0 1.09
Bruneau[26] SC-20 2.5 420.6 46.8 38.1 8.25 0 1.09
Alzeni and SC-21 BE-SC 2.76 427.5 48.8 9.5 11.1 0 1.15
Bruneau[26] walls
Nie et al.[27] SC-22 2 310 95 12.3 7.1 24.3 1.73
SC-23 2 310 90 12.3 7.1 25.2 1.74
SC-24 2 310 90 12.3 7.1 25.2 1.85
SC-25 2 360 95 14.9 5.85 24.8 1.55
SC-26 2 450 94 18.9 4.6 24.4 1.54
SC-27 2 310 70 12.3 7.1 24.5 1.78
SC-28 2 310 105 12.3 7.1 24.1 1.96
SC-29 2 360 95 14.9 7.1 24.9 1.58
SC-30 2 310 95 12.3 7.1 24.3 1.8
SC-31 2 450 87 12.1 7.04 26.1 1.94
ASGARPOOR ET AL.
TABLE 4 (Continued)

Specimen Aspect Steel yield strength Concrete compressive strength Slenderness Reinforcement Axial load ratio Peak strength
Researcher ID Shape ratio (MPa) (MPa) ratio ratio (%) (%) ratio
SC-32 1.5 450 85 12.1 7.04 26.2 1.7
ASGARPOOR ET AL.

SC-33 1 450 94 12.1 7.04 25.2 1.36


[31]
Luo et al. SC-34 1.8 276.8 35.5 18.7 15.1 40 1.01
SC-35 1.8 276.8 35.5 18.7 15.1 25 0.61
SC-36 1.8 276.8 35.5 18.7 15.1 40 0.95
SC-37 1.8 276.8 35.5 18.7 14.7 40 0.76
SC-38 2.5 276.8 35.5 18.7 15.1 40 0.94
SC-39 2.5 336.2 35.5 25 11.5 25 0.88
SC-40 1.8 336.2 35.5 25 11.5 25 0.78
SC-41 1.8 276.8 51.3 18.7 15.1 25 1.08
[32]
Eom et al. SC-42 FL-SC 3.85 383 39.7 15 19.4 0 1.38
walls
19 of 22
20 of 22 ASGARPOOR ET AL.

for the prediction of the moment capacity of SC walls without boundary elements causes acceptable results in low aspect ratio walls, including
Kurt et al.,[50] Cheng and Zhou,[51] and Epackachi et al.[25] Hence, the accuracy of the results highly depends on the aspect ratio.
Based on the provisions of chapter H-7 of AISC 341-16,[49] the nominal moment capacity of SC walls with boundary elements is equal to its
plastic moment capacity, Mp, assuming all concrete in compression has reached its compressive strength, f 0c , and considering the steel in tension
and compression has reached its yield strength, Fy. One of the missing parameters in the AISC equation is the axial load ratio. Table 4 presents
that the shear strengths of SC walls with a high axial load ratio have meaningful discrepancies with code results. Although the use of a large slen-
derness ratio, S/ts, leads to early buckling of steel faceplates and noncompactness of the section, there is no limitation for this parameter in AISC
341-16. In most of the SC wall experiments, the slenderness ratio was designed conservatively. The effect of changing this ratio or eliminating
the connectors on the response of several SC walls is investigated in Section 5.3 of this paper. The study by Zhang et al.[37] provides a suitable
guideline to design the spacing of tie bolts in SC walls. Another important parameter that the code ignores to consider is the range of concrete
compressive strength that can be used. Nie et al.[27] have used high-strength concrete in their tested SC walls, and AISC 341-16 formulation
underestimates the peak strength of their walls. A large reinforcement ratio leads to a brittle failure in SC walls, so the wall cannot reach its plastic
moment capacity, as observed in the walls of Luo et al.[31] in Table 4.
Based on the results of experimental studies and effects of various design variables on the peak strength of SC walls, it is concluded that the
use of the PSD method for peak strength prediction of SC walls with boundary elements leads to low accuracy if the walls are squat, with consid-
erable compressive axial load ratio, high concrete compressive strength (more than 50 MPa), large slenderness ratio (more than 40), and large rein-
forcement ratio (more than 10%). As a result, for an accurate calculation of the peak strength of SC walls, all key design variables shall be
considered.

6.2 | Plastic hinge length

Macro models are used to represent the overall behavior of the structural element and known as an efficient and practical tool that does not
require high computational efforts making it suitable for engineering applications. The accuracy of the macro models significantly depends on the
simplifying assumptions and the idealized material models upon which the model is developed. One of the important assumptions in macro model-
ing of the flexural-controlled shear walls is the length of the plastic hinge, where nonlinear materials are used to simulate the overall nonlinear
response of the shear wall element. To reduce the computational cost of analysis and element convergence problems in nonlinear analysis of full-
scale buildings, including RC or SC shear walls, elastic material is used for elements beyond the region expected to exhibit nonlinear response.
The response of elastic parts does not depend on the mesh size of the elements, and no failure criterion is considered. These simplifying assump-
tions significantly reduce the time of analyses in large-scale finite element analyses.

TABLE 5 Plastic hinge length of numerical models

Specimen ID Shape Wall aspect ratio Wall height, h (mm) Plastic hinge length, lp (mm) Plastic hinge ratio, lp/h
Wall-1 DS-SC walls 1 1525 457 0.3
Wall-2 DSE-SC walls 2.45 3050 1524 0.5
Wall-3 BE-SC walls 2.56 3050 1524 0.5
Wall-4 BE-SC walls 1.98 2560 1270 0.5
Wall-5 BE-SC walls 1.5 1120 635 0.56
Wall-6 DS-SC walls 2.37 950 305 0.32
Wall-7 DSE-SC walls 3 2740 965 0.35
Wall-8 DSE-SC walls 3 2740 1372 0.5
Wall-9 DS-SC walls 1.78 1420 737 0.51
Wall-10 DSE-SC walls 1.78 1420 584 0.41
Wall-11 BE-SC walls 1.8 2160 1270 0.58
Wall-12 BE-SC walls 2.5 3000 1651 0.55
Wall-13 DSE-SC walls 3.86 3850 1422 0.36
Wall-14 DSE-SC walls 3.86 3850 2184 0.56
Wall-15 DSE-SC walls 3.86 3850 1219 0.31
Wall-16 DSE-SC walls 2.66 4000 1270 0.32
Wall-17 FL-SC walls 2.66 4000 1270 0.32
ASGARPOOR ET AL. 21 of 22

To date, the length of the plastic hinge at the base of SC shear walls has not been investigated systematically. In the 17 finite element-based
analyses conducted in this study, the length of the plastic hinge is calculated using the concept of yield strength capacity provided in AISC
341-16, where the plastic hinge region is defined as a zone from the base of the wall up to a height where the steel reaches to its minimum yield
strength, Fy, and concrete reaches to 0:7f 0c . The height of the plastic hinge region is called plastic hinge length, lp. Table 5 provides the cross-
sectional shape of SC walls, wall aspect ratio, wall height, plastic hinge length, and the ratio of plastic hinge length to the wall height for all SC wall
models. The studied walls include four different cross-sectional shapes, and their aspect ratio is larger than 1.0. Based on the analysis results, the
plastic hinge length of the SC walls in the considered domain is in the range of 0.3 to 0.6 of the wall height implying at least 40% of the wall height
from the top can be modeled using elastic material with no significant effect on the wall response. As seen in Table 5, the length of the plastic
hinge in SC walls with boundary elements is more than that of walls without boundary elements. Since the plastic hinge length in SC walls
depends on several design parameters such as the shape of the wall cross-section, wall aspect ratio, material properties, and axial load ratio, devel-
opment of a predictive equation for the calculation of the plastic hinge length in SC walls, requires a larger number of numerical models which can
be investigated as further studies.

7 | C O N CL U S I O N

In this paper, comprehensive numerical analyses were conducted to simulate the cyclic responses of 17 SC walls, tested under quasi-static cyclic
loading. The SC walls were differed by their section geometry and axial loading condition, including rectangular, T-shaped, and flanged walls, iso-
lated and coupled walls, with and without boundary elements, endplates, and axial compressive load. The Winfrith concrete and plastic kinematic
steel models were used to model concrete and steel, respectively. The 8-noded solid infill concrete and 4-noded shell steel faceplate elements
were used in the models. The connectors, including shear studs and tie rods, were modeled by beam elements with the same material model as
the steel faceplates. The results of this study are summarized as follows:

• The force-displacement relationships of numerical models matched well with their corresponding test diagrams, resulting in acceptable consis-
tency between the initial stiffness, peak strength, unloading/reloading stiffness, and post-peak strength degradation of the test and analysis
results. Despite some discrepancies between the experimental measurements and numerical predictions, the models could capture the global
pinching behavior of SC wall models appropriately. Besides, the LS-DYNA predicted damage to SC walls, including concrete cracking, faceplate
buckling, concrete crushing, and faceplate fracture, were in good agreement with test observations.
• The global response of SC walls was not significantly affected by varying the friction coefficient between the infill concrete and steel face-
plates from 0.3 to 1.
• Modeling the bond-slip between headed studs and concrete showed no significant changes in the studied SC wall models.
• The effect of different configurations of studs and tie rods on SC wall response was investigated by varying the slenderness ratio and eliminat-
ing some of the connectors in SC walls. The importance of the configuration of connectors is dominant in regions where plastic hinges form,
which occurs mostly at the wall base. The extent of influence of S/ts value on the global response of SC walls depends on the wall geometry
(the existence of boundary elements, batten plates, and endplates) and loading condition (axial compressive load). Using thicker steel face-
plates, the presence of axial compressive load and applying stiffening elements decrease the sensitivity of SC wall strength and ductility to the
S/ts value. The studs with adequate embedment length can behave like tie rods and prevent the buckling of faceplates by attaching them to
concrete.
• The most common approach for predicting the peak lateral strength of SC walls is using the plastic stress distribution method, which is rec-
ommended by AISC 341-16. Based on the results of several finite element analyses, the most important design variables with a significant
effect on the strength and performance of SC walls under seismic loading are wall aspect ratio, steel yield strength, concrete compressive
strength, slenderness ratio, reinforcement ratio, and axial load ratio. Therefore, it is worth improving the PSD method to consider the effects
of key design variables in predicting the peak lateral strength of SC walls.
• Based on the results of the analyses of SC walls with different design variables, the plastic hinge in SC walls, in which their behavior is domi-
nated by flexure, forms at 0.3 to 0.6 of the wall height from the base of the wall. This range of plastic hinge length is recommended to be used
for macro modeling of SC walls in future studies.

DATA AVAI LAB ILITY S TATEMENT


Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.

ORCID
Seyed Rasoul Mirghaderi https://orcid.org/0000-0003-4654-0689
22 of 22 ASGARPOOR ET AL.

RE FE R ENC E S
[1] N. Sasaki, H. Akiyama, M. Narikawa, K. Hara, M. Takeuchi, S. Usami, Study on a concrete filled steel structure for nuclear power plants (part 3). Shear
and bending loading tests on wall member, 1995.
[2] S. Usami, H. Akiyama, M. Narikawa, K. Hara, M. Takeuchi, N. Sasaki, Study on a concrete filled steel structure for nuclear power plants (part 2).
Compressive loading tests on wall members, 1995.
[3] T. Takeda, T. Yamaguchi, T. Nakayama, K. Akiyama, Y. Kato, Experimental study on shear characteristics of a concrete filled steel plate wall, 1995.
[4] M. Ozaki, S. Akita, H. Osuga, T. Nakayama, N. Adachi, Nucl. Eng. Des. 2004, 228(1–3), 225.
[5] N. Suzuki, H Akiyama, M Narikawa, K Hara, M Takeuchi, I Matsuo, Study on a concrete filled steel structure for nuclear power plants (part 4).
Analytical method to estimate shear strength, 1995.
[6] F. J. Vecchio, I. McQuade, Nucl. Eng. Des. 2011, 241(8), 2629.
[7] M. H. Kisa, S. B. Yuksel, N. Caglar, J. Build. Eng. 2020, 33, 101570.
[8] C. Xiao, A. Zhu, J. Li, Y. Li, J. Build. Eng. 2020, 34, 101909.
[9] Y. Hu, J. Zhao, D. Zhang, Y. Li, J. Build. Eng. 2020, 31, 101376.
[10] J.-B. Yan, A. Chen, T. Wang, Structure 2020, 28, 407.
[11] Q. Guo, P. Zhang, Y. Fu, F. Ke, Structure 2020, 27, 1570.
[12] J. Liew, Y. Chua, Z. Dai, Structure 2019, 21, 135.
[13] A. Ali, D. Kim, S. G. Cho, Nucl. Eng. Technol. 2013, 45(1), 89.
[14] S. Rafiei, K. M. A. Hossain, M. Lachemi, K. Behdinan, M. S. Anwar, Eng. Struct. 2013, 56, 46.
[15] E. Kurt, A. Whittaker, A. Varma, P. Booth, SC wall piers and basemat connections: numerical investigation of behavior and design, 2013.
[16] S. Epackachi, A. S. Whittaker, A. H. Varma, E. G. Kurt, Eng. Struct. 2015, 100, 369.
[17] E. Polat, M. Bruneau, Eng. Struct. 2017, 148, 63.
[18] S. Dey, A. K. Bhowmick, Structure 2016, 6, 59.
[19] N. Haghi, S. Epackachi, M. T. Kazemi, Structure 2020, 23, 383.
[20] J. Li, F. Li, C. Liu, J. Miao, J. Build. Eng. 2020, 31, 101359.
[21] A. Plumier, D. Dragan, N. Q. Huy, H. Degée, Structure 2017, 9, 185.
[22] N. Haghi, S. Epackachi, M. T. Kazemi, Structure 2020, 27, 102.
[23] S. Epackachi, A. S. Whittaker, A. Aref, Eng. Struct. 2017, 133, 105.
[24] E. Polat, Uludag Univ. J. Faculty Eng. 2020, 25(1), 139.
[25] S. Epackachi, N. H. Nguyen, E. G. Kurt, A. S. Whittaker, A. H. Varma, J. Struct. Eng. 2015, 141(7), 04014176.
[26] Y. Alzeni, M. Bruneau, Tech. Rep. MCEER 2014, 14, 9.
[27] J.-G. Nie, H. S. Hu, J. S. Fan, M. X. Tao, S. Y. Li, F. J. Liu, J. Constr. Steel Res. 2013, 88, 206.
[28] S. G. Cho, W. K. Park, G. H. So, S. T. Yi, D. Kim, KSCE J. Civil Eng. 2015, 19(3), 698.
[29] M. Bruneau, Eng. Struct. 1998, 20(12), 1063.
[30] Q. Ge, T. He, F. Xiong, P. Zhao, Y. Lu, Y. Liu, N. Zhou, J. Civil Eng. Manag. 2020, 26(3), 227.
[31] Y. Luo, X. Guo, J. Li, Z. Xiong, L. Meng, N. Dong, J. Zhang, Adv. Struct. Eng. 2015, 18(11), 1845.
[32] T.-S. Eom, H. G. Park, C. H. Lee, J. H. Kim, I. H. Chang, J. Struct. Eng. 2009, 135(10), 1239.
[33] LS-DYNA keyword user's manual, Livermore Software Technology Corporation, Vol. I, 7374, 2007, 354.
[34] M. Asgarpoor, A. Gharavi, S. Epackachi, Structure 2021, 29, 1322.
[35] N. S. Ottosen, J. Eng. Mech. Div. 1977, 103(4), 527.
[36] CEB-FIP, Comite Euro-International du beton, Model code, 2010.
[37] L. Schwer, The Winfrith concrete model: beauty or beast? Insights into the Winfrith concrete model, 8th European LS-DYNA Users Conference,
2011; 23–24.
[38] B. Broadhouse, DRASTIC: a computer code for dynamic analysis of stress transients in reinforced concrete, 1986.
[39] LS-DYNA keyword user's manual, Livermore Software Technology Corporation, vol. II, 2013.
[40] B. Rabbat, H. Russell, J. Struct. Eng. 1985, 111(3), 505.
[41] A.K. Chopra, Dynamics of structures. 1995.
[42] B. Evirgen, A. Tuncan, EUROSTEEL 2014, 12, 419.
[43] J. McCormick, T. Nagae, M. Ikenaga, P. C. Zhang, M. Katsuo, M. Nakashima, Earthq. Eng. Struct. Dyn. 2009, 38(12), 1401.
[44] T. Nagae, M. Ikenaga, M. Nakashima, K. Suita, J. Struct. Constr. Eng. 2006, 71(606), 217.
[45] N. S. Anderson, D. F. Meinheit, PCI J. 2000, 45(5), 46.
[46] J. G. Ollgaard, R. G. Slutter, J. W. Fisher, AISC Eng. J. 1971, 8(2), 55.
[47] C.-S. Shim, P.-G. Lee, T.-Y. Yoon, Eng. Struct. 2004, 26(12), 1853.
[48] K. Zhang, A. H. Varma, S. R. Malushte, S. Gallocher, Nucl. Eng. Des. 2014, 269, 231.
[49] American Institute of Steel Construction, Seismic provisions for structural steel buildings (AISC 341-16), 2016.
[50] E. G. Kurt, A. H. Varma, P. Booth, A. S. Whittaker, J. Struct. Eng. 2016, 142(6), 04016026.
[51] C. Cheng, D. Zhou, Experimental study on seismic behavior of composite concrete and double-steel-plate shear walls with binding bars, 6th Interna-
tional Conference on Advances in Experimental Structural Engineering & 11th International Workshop on Advanced Smart Materials and Smart Struc-
tures Technology. August 1–2, University of Illinois Urbana-Champaign, United States, 2015.

How to cite this article: M. Asgarpoor, A. Gharavi, S. Epackachi, S. R. Mirghaderi, Struct Design Tall Spec Build 2021, e1863. https://doi.
org/10.1002/tal.1863

View publication stats

You might also like