You are on page 1of 25

Received: 4 May 2017 Revised: 15 August 2017 Accepted: 7 October 2017

DOI: 10.1002/tal.1447

RESEARCH ARTICLE

Strength and post‐yield behavior of T‐section steel encased by


structural concrete
J.D. Nzabonimpa | Won‐Kee Hong | Jisoon Kim

Department of Architectural Engineering,


Kyung Hee University, Yongin, Republic of Summary
Korea In this study, the seismic performance of unsymmetrical steel–concrete composite precast beams
Correspondence with T‐shaped steel section were numerically explored and validated by their earlier experimental
Won‐Kee Hong, Department of Architectural investigation. This design is based on the proposed calibrated finite element model in which key
Engineering, Kyung Hee University, Yongin
446‐701, Republic of Korea.
damage parameters for the evaluation of the nonlinear, post‐yield behavior of the precast
Email: hongwk@khu.ac.kr composite steel beams were identified. The proposed nonlinear finite‐element‐based numerical
Funding information model uses various parameters, including the dilatation angle and concrete‐damaged plasticity,
Basic Science Research Program through the to simulate the nonlinear behavior of unsymmetrical composite precast beams with T‐section
National Research Foundation of Korea (NRF);
steel. Greater seismic capacity with greater ductility, contributing to a maximized structural
Ministry of Education, Grant/Award Number:
2016R1D1A1A02937558 capacity within the composite precast beams was introduced by the effective use of the 2
materials, steel and concrete, and shown by the nonlinear hysteretic investigation of unsymmet-
rical steel–concrete composite precast beams that was validated experimentally. The post‐yield
structural capacity found via the numerical analysis agrees with experimental results when the
concrete‐damaged plasticity of the plastic‐damaged seismic model for concrete and the von
Mises criteria of the steel section were introduced into the finite element model. Practical design
parameters and recommendations were eventually suggested by examining the influence of
precast composite beams with unsymmetrical steel sections on the concrete degradations and
damage evolution.

KEY W ORDS

damaged plasticity, dilation angle, fracture energy, reinforced concrete precast frames, steel–
concrete composite precast frames, yield surface

1 | I N T RO D U CT I O N

1.1 | Limitation of the finite element analyses (FEA) of beams with concrete‐damaged plasticity
Current literature analyses of beams based on damaged plasticity are primarily limited to reinforced concrete beams, because precast composite
steel beams are less commonly used in the construction industry. Considerable research, including the work by Prakash et al.,[1] focuses on defining
the plastic parameters of three‐dimensional (3D) finite element (FE) models of stud‐connected steel–concrete composite girders. In their study, the
post‐yield behavior of composite girders having either full or partial shear connection with concrete slabs was investigated. Wahalathantri et al.[2]
present a material model to simulate load‐induced cracking in reinforced concrete elements using an ABAQUS FE package. Two numerical material
models were used and combined to simulate the complete stress–strain behavior of concrete under compression and tension, including damage
properties. Grassl and Jirásek[3] concluded that the combination of plasticity, based on the effective stress, and the isotropic damage, driven by
plastic strain, are suitable for predicting the failure of concrete in a wide range of loading cases from uniaxial tension to triaxial compression in their
study, “Damage‐plastic model for concrete failure.” Numerical simulations concerning uniaxial and biaxial compression and the uniaxial tension of a
sample concrete specimen was developed by Szczecina and Winnicki,[4] and recommends a viscosity parameter of 0.0001 and a dilation angle of 5°.
Kulkarni et al.[5] present a nonlinear FEA of hybrid‐steel concrete connections, identifying critical parameters influencing the joint's behavior, such
as the axial load on the column, the thickness of the connection plate, and the continuation of beam bottom reinforcement. Korkmaz and Tankut[6]

Struct Design Tall Spec Build. 2018;27:e1447. wileyonlinelibrary.com/journal/tal Copyright © 2017 John Wiley & Sons, Ltd. 1 of 25
https://doi.org/10.1002/tal.1447
2 of 25 NZABONIMPA ET AL.

also investigate the performance of a precast concrete beam‐to‐beam connection subject to reversed cyclic loading to observe and investigate the
seismic behavior of the connection detail proposed by the collaborating company in order to develop a “moment resisting precast concrete
beam‐to‐beam connection.” Genikomsou and Polak[7] present a damaged plasticity model in FEA using ABAQUS by comparing the experimental
results of punching shear testis on concrete slabs based on force‐displacement response to their numerical results obtained from the static and
quasi‐static force‐displacement response. In their study, two viscosity parameters of 0.00001 and 0.0005 are used in the static analysis. Omidi
and Lotfi[8] performed a four‐point bending test to demonstrate the capabilities of the hyperbolic potential function to treat the singularity of
the original linear form of plastic flow proposed by Lee and Fenves[9] in real 3D applications. In addition, Hawileh et al.[10] developed a detailed
3D nonlinear FE model to study the response and predict the behavior of precast hybrid beam–column connections subjected to cyclic loads. These
were tested at the National Institute of Standards and Technology laboratory, successfully providing a practical and economical tool to investigate
the behavior of such connections. Malm[11] summarized information relating to damaged plasticity models with many efficient examples; this work
provided this study with useful assistance. He also evaluated both methods in design codes based on experiments and advanced FEA with a
nonlinear material description. Additional studies were also undertaken to comprehend the behavior of structural steel for composite members.
Mises and Hencky[12] further developed the yield criterion of steel sections for composite members that were originally proposed by Huber in
1904. It was suggested that the yielding occurs when the distortion energy is equivalent to the yield stress in a simple tensile test (Ugural and
Fenster[13]). In this paper, the von Mises yield criterion is considered in FE models to define the yield criterion of steel sections and rebars for
the proposed composite beam.
A significant number of studies were conducted to assess the structural performance of precast steel–concrete beams and columns subjected
to seismic loadings.[14–20] Some of these studies highlighted the advantages of the application of steel–concrete composite members, and they
concluded that steel–concrete structures exhibit large flexural strength and high ductility under cyclic loadings. Throughout experimental investi-
gations, Kindmann et al.[16] contributed to the design of steel–concrete composite beams. In their study, a total of 13 test specimens were loaded
to failure to uncover the importance of reinforced concrete between the flanges of the steel for the calculation of the flexural strength and the
deflection of steel–concrete composite beams. Useful recommendations were provided for the modification of the Eurocode 4, which was found
to be too conservative for the calculation of load bearing capacity and deflection. Another study performed by Ricles and Paboojian[15] performed a
parametric study on the structural behavior of steel–concrete beams. They studied the degree of concrete confinement required to attain enough
ductility under seismic loadings. Their findings indicated that concrete encased steel members demonstrated exceptional structural strength with
high energy dissipation. Similarly, Dundar et al.[20] studied the confinement effect of steel‐composite columns via experimental and numerical tests.
In their work, the ultimate strength capacities obtained from various concrete models were compared with each other, and the load‐deflection
behavior for steel–concrete composite beam was computed.
Our study is devoted to the evaluation of the seismic performance of unsymmetrical steel–concrete composite precast beams with T‐section
steel. The proposed steel–concrete beam is believed to contribute to the fast construction of the precast composite steel frames, and it is proposed
to replace conventional cast‐in‐place frames because these traditional construction methods are expensive and they lengthen the construction
time. Many studies regarding the flexural strength of steel–concrete composite beams presented previously did not recommend practical FEA
parameters and recommendations for the inelastic response and behavior of composite beams. In this regard, this paper investigated the ductility
with confining pressures based on dilation angles and stiffness degradation of the proposed steel–concrete beams. The nonlinear, inelastic
response and behavior of the proposed precast composite beams subjected to cyclic loads were calculated based on the suggested nonlinear FE
parameters. Practical FEA parameters and recommendations were eventually suggested by examining the influence of precast composite beams
with unsymmetrical steel sections on the concrete degradations and damage evolution. The von Mises yield criterion is considered in FE models
to define the yield criterion of steel sections and rebars for the proposed composite beam.

1.2 | Research significance of this study


In their previous paper, the authors proposed wide‐flange steel beams in which the bottom flange is encased in precast concrete as a replacement
for conventional cast‐in‐place composite beams, as shown in Figure 1. The effective interaction between these two materials can reduce the size of
the steel beams. Successful construction was possible because the slabs were constructed on top of the edges of the U‐shaped precast concrete
(Figure 1 (a)) instead of on top of the steel flanges, reducing the thickness of the slab and beam and leading to a significant reduction in floor height.
The bottom steel reinforcement and concrete are pre‐integrated with the bottom flange of the steel beam at the manufacturing plant (Figure 1 (b)).
The concrete covering steel reinforcement and flange are cast at the traditional construction site as shown in Figure 1 (b,c) in which the proposed
precast composite beams are compared with the conventional cast‐in‐place construction of steel–concrete beams requiring pour forms and
concrete casts, creating significant delays and safety issues for the sites. In a previous application of these composite precast beams, shown in
Figure 2 (b), to an 18‐story building with a total height of 67.64 m, the building was able to be redesigned and built with 19 stories with a total
height of 66.4 m while maintaining the same floor–ceiling height. The average reduction in floor height per floor was 220 mm. Figure 2 shows
the transport of the manufactured products and the erection of the composite precast beams. In their previous study, the authors developed an
analytical model based on strain compatibility to calculate the post‐yield behavior of the composite beams shown in Figure 3; however, concrete
plasticity in the composite beams was not taken into consideration. The objective of this research is to evaluate the seismic performance of
unsymmetrical steel–concrete composite precast beams with T‐section steel based on the concrete‐damaged plasticity. The nonlinear, inelastic
NZABONIMPA ET AL. 3 of 25

(a) Slabs being constructed on top of the edges of the U-shaped


precast concrete (Hong et al. [14])

(b) Concrete pre-integrated with the bottom flange of the steel beam
at the manufacturing plantreplacing conventional cast-in-place concrete/
[14]
composite frames(Hong et al. )

FIGURE 1 Concrete/steel–concrete
composite precast frames versus cast‐in‐place
frames. (a) Slabs being constructed on top of
the edges of the U‐shaped precast concrete
(Hong et al.[21]), (b) concrete pre‐integrated
with the bottom flange of the steel beam at
the manufacturing plant replacing
conventional cast‐in‐place concrete/
composite frames (Hong et al.[21]), and (c)
conventional cast‐in‐place construction (c) Conventional cast-in-place construction

response and behavior of the proposed precast composite beams subjected to cyclic loads was calculated based on the suggested nonlinear FE
parameters; the numerical results of the seismic performance were validated via test data measured in their earlier study.[22]

2 | M A T E RI A L P R O P ER T I ES

2.1 | Concrete
2.1.1 | Elastic‐plastic model for concrete and flow rule
Dilation caused by plastic distortion can be reproduced by the plastic potential function G, by Lubliner et al.[23] The evolution of the inelastic
∂GðσÞ
displacement in the fracture process zone is determined by the flow rule defined as ε_pl ¼ λ_ . Plastic flow develops along the normal to the
∂σ
4 of 25 NZABONIMPA ET AL.

(a) Transport of the manufactured products (b) Erection of the pre-cast composite beams

FIGURE 2 Application to a 19‐story building; (a) transport of the manufactured products and (b) erection of the precast composite beams

FIGURE 3 T‐section steel encased in precast concrete

plastic potential, and not to the yield surface for a non‐associated flow (Galvez et al.[24]). The non‐associated plastic potential function reaches the
asymptotic line of the linear Drucker–Prager flow potential at high confining pressure stress, which intersects the hydrostatic pressure axis at 90o,
not intersecting the yield surface, as shown in Figure 4. The derivative of the plastic flow potential function with respect to the effective stress,
which determines the direction of the rate of plastic strain (ε_pl ), defines the rate of plastic strain (ε_pl ). The length of the strain increment vector,
ε_pl , is governed by a scalar hardening parameter denoted as λ;
_ which varies depending on the straining process (Chen and Han[25]).

2.1.2 | Dilation angle and eccentricity


Equation 1 describes the Drucker–Prager hyperbolic plastic potential function used in the damaged plasticity concrete model depicted in Figure 4,

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
GðσÞ ¼ ðϵσt0 tan ψÞ2 þ q2 −ptanψ; (1)

FIGURE 4 Non‐associated plastic flow potential, G (Hibbit et al.[26])


NZABONIMPA ET AL. 5 of 25

where ψ and σto are the dilation angle measured in the p‐q plane at high confining pressure and the uniaxial tensile stress, respectively, p
is the hydrostatic pressure and q is the Mises equivalent effective stress. Considerable volume change induced by plastic distortion, termed
dilatancy, is caused by inelastic strains in a brittle material such as concrete. The dilatancy in a concrete‐damaged plasticity model is
modeled using the dilation angle. ABAQUS uses the Drucker–Prager hyperbolic plastic potential function for a damaged plasticity concrete
model (Hibbit et al.[26]) that defines dilatancy via the dilation angle, as demonstrated in Figure 4. Three types of models, linear, hyperbolic,
and exponential, were suggested. In this study, the concrete‐damaged plasticity model was constructed based on a non‐associated
hyperbolic function to reproduce the change in volume. A scalar hardening parameter in the illustration indicates the length of the strain
increment vector.
The inclination of the plastic potential at high confining pressure stresses determined by the dilation angles governs the direction of the plastic
potential. In the non‐associated flow rule, the plastic strain vector is normal to the plastic potential function, and not to the yield surface. The use of
low values for the dilation angles causes brittle behavior, while ductile behavior will be produced at higher values. Malm[27] suggested a dilation
angle between 30o and 40o for the best agreement with the experimental responses obtained from reinforced concrete tests. The dilation angle
for steel–concrete precast composite beams was sought based on the parameter sensitivity analysis.

2.1.3 | Viscoplastic regularization


The ABAQUS[28] regularizes the concrete‐damaged plasticity model by using viscoplasticity to permit stresses to be outside of the yield surface.
The viscoplastic strain rate tensor is defined in Equation 2,

1 pl pl 
_ε pl
v ¼ ε −εv ; (2)
μ

which was proposed by Duvaut‐Lions. Here, μ is the viscosity parameter representing the relaxation time of the viscoplastic system under
consideration and the plastic strain evaluated in the inviscid model. The plastic flow rate is low when viscosity (μ) increases, which decreases
not only the plastic degradation but also the rate of plastic fracture of the section, enhancing the convergence of the calculation. The viscosity
was also applied to the degradation variable for the viscoplastic system, as shown in Equation 3. The viscosity parameter μ can control the rate
of the plastic fracture to enhance the convergence of the numerical process.
The viscous stiffness degradation variable for the viscoplastic system to control the rate of degradation is defined as

1
d_ v ¼ ðd−dv Þ; (3)
μ

where d is the degradation variable for the viscoplastic system. The damage variable considering viscosity parameter μ is defined as dv. Therefore,
the stress–strain relationship between the viscoplastic model and the viscous stiffness degradation variable is obtained using Equation 4:

 
σ ¼ ð1−dv ÞDel0 : ε−εpl
v : (4)

The rate of convergence of the model in the softening region can be improved via viscoplastic regularization with a small value relative to the
time increment without compromising the analysis results. In this study, a change in viscosity does not significantly influence the analysis results of
the specimen. Fenves proposed a viscosity parameter with a value that is 15% of time increment step in order to overcome severe convergence
difficulties for highly weakened structural systems. A default value was used for the viscosity parameter, because no severe convergence
difficulties occurred in the investigation of the structural behavior of the steel–concrete precast composite beam.

2.2 | Damaged plasticity model


2.2.1 | Damaged plasticity model for concrete with constitutive relation; uniaxial tension and compression stress behavior
The failure mechanics of concrete in tension and compression for low levels of confinement is associated with both stiffness degradation and
inelastic deformations characterized by a coupling between damage and plasticity, as seen in Equation 5. It is assumed that the strain rates can
be decomposed as

ε_ ¼ ε_ el þ ε_ pl ; (5)

where ε_ is the total strain rate, ε_ el is the elastic strain rate, and ε_ pl is the plastic strain rate. The stress–strain relationships are written as indicated in
Equation 6,

 
σ ¼ ð1−dÞDel0 : ε−εpl el
v ¼ D : ε−ε ;
pl
(6)

based on scalar‐damaged plasticity, where Del0 is the undamaged elastic stiffness of the material, Del ¼ ð1−dÞDel0 is the degraded elastic stiffness, and
d is the scalar stiffness degradation variable. When the unloading response with the degraded elastic stiffness of the concrete is taken into account,
6 of 25 NZABONIMPA ET AL.

the stress–strain relationship under uniaxial tension and compression loading become, respectively (Equations 7 and 8),

 
σt ¼ ð1−dt ÞE0 εt −ε_ tpl ; (7)

 
σc ¼ ð1−dc ÞE0 εc −ε_ cpl ; (8)

where E0 is the undamaged elastic stiffness of the material. The two damage variables, dt and dc, which are functions of plastic strain, govern the
damage response of the concrete under tension and compression. The degradation of elastic stiffness differs significantly between tension and
compression, and its effects become more pronounced as the plastic strain increases. Equation 9 describes stress states representing the stiffness
recovery effects associated with stress reversals based on St and Sc,

ð1−dÞ ¼ ð1−st dc Þð1−sc dt Þ; 0 ≤ st ; sc ≤ 1: (9)

2.3 | Concrete damaged plasticity


Concrete is assumed to fail as a result of tensile cracking and compressive crushing under low confining pressures based on concrete damaged
plasticity. However, the brittle behavior of concrete becomes insignificant when the confining pressure is large enough to prevent cracking
(Chaudhari and Chakrabarti[29]). In our model, both the compressive and tensile behaviors of concrete are defined via the concrete‐damaged
plasticity model, and concrete is discretized to guarantee accurate results, similar to concrete compressive strength. The Kent‐Park[30] constitutive
relationship with calibrated data to 33.1 MPa for concrete is given in Figure 15 (a,b).

3 | FE MODEL

3.1 | Use of C3D8R


Two types of 3D solid elements were used in the modeling of the embedded T‐steel section, concrete beam, and reinforcing bars. A total of 56,937
elements were modeled as hexahedral elements of type C3D8R, while the remaining 96 elements were assigned as quadrilateral elements of type
R3D4. Reduced‐integration elements (C3D8R) were used to reduce the running time for the executed ABAQUS analyses. These elements
demonstrate characteristics similar to that of C3D8‐type elements, except that 8 nodes are reduced to a single integration point, meaning that
the number of degrees of freedom is drastically reduced in C3D8R elements, as depicted in Figure 5. ABAQUS[28] uses lower‐order integration
to evaluate the element stiffness matrix for reduced‐integration elements.
Elements of type R3D4 were used to discretize the rigid body with a reference point on which a prescribed lateral displacement was applied, as
indicated in Figure 6. Each of these elements has 4 nodes, each with 3 degrees of freedom. Figure 6 demonstrates how the element connectivity is
maintained via the right‐hand rule by going around the nodes of elements to define a positive normal for R3D4.

3.2 | Modeling of embedded elements


The main steel reinforcements are modeled using 3D elements of type C3D8R. The von Mises yield criterion with isotropic strain hardening is used
to describe the constitutive behavior of the reinforcing bars. Similarly, the bolted T‐steel section and rebars are modeled as elements of type
C3D8R. All embedded elements are embedded into concrete (host element) by defining the embedded region provided by ABAQUS. As shown
in Figures 9 and 14 and Table 1, two headed studs of 22 mm diameter spaced at 95 mm on the web and one spaced at 450 mm on the flange were
installed on the T‐steel section to prevent slippages from occurring between the embedded elements and concrete during the analysis. A perfect
bond between the embedded elements and concrete is realized so that no slippage can occur during the analysis. Both the embedded and host
elements for the steel–concrete composite beam are demonstrated in Figure 7. The tensile yield strength of longitudinal reinforcement was tested
to be 437.5 MPa, and the stirrups for the composite beam were modeled with 400 MPa. Steel plates (SM490) with an initial tensile yield strength of
325 MPa and a Poisson's ratio of 0.3 were accounted for in the FEA. The density and Young's modulus for the steel plates were 7.85 × 10−09ton/
mm3 and 205,000 MPa, respectively, whereas the Young's modulus for rebars was 206,000 MPa. High‐strength bolts (F10 T) with a yield strength
of 900 MPa and a Poisson's ratio and Young's modulus similar to those of steel plates were also used in the analysis. The typical stress–strain curve
of the rebars is assumed to be elasto‐plastic and identical in both tension and compression. Although the 2D model assumption of treating the plate
area as equivalent to the truss element was approximate, solid, plane stress, and truss elements were used for the concrete, steel plates, and
reinforcing bars, respectively, in this study.
NZABONIMPA ET AL. 7 of 25

One integration point


8 7

5 6
1
4 3

1 2
FIGURE 5 Eight‐node brick element with reduced integration (C3D8R)

4 3 Face SPOS
n

n
1 2
Z
Y 1
X face SNEG 2

FIGURE 6 Positive normal for R3D4 elements (ABAQUS)

TABLE 1 Description of the three identical test specimens (Hong et al.[22])


Stirrup Headed stud bolt (l = 65 mm)
Web Bottom flange
Spacing(mm) Diameter(mm)
Category Spacing (mm) Diameter(mm) Spacing(mm) Diameter(mm)

Specimen PS1 500 10 95 13 450 13


Specimen PS2 500 10 95 13 450 13
Specimen PS3 500 10 95 13 450 13

FIGURE 7 Defining constraint: Embedded region


8 of 25 NZABONIMPA ET AL.

4 | EXPERIMENTAL INVESTIGATION AND RESULTS

4.1 | Design and fabrication of specimens


Three test specimens were manufactured for testing in their earlier study[22] in order to estimate the seismic performance of unsymmetrical
steel–concrete composite precast beams with T‐section steel, as shown in Figures 8 and 9 (a). The composite beams were 300 mm wide with a
depth of 350 mm and a steel section, T‐150 × 150 × 6.5 × 9, with a yield strength of 330 MPa. The reinforcing steel bars are 2‐HD13 and
4‐HD22, which were placed at the bottom and top reinforcement, respectively. Average values of material properties were obtained from the
tensile coupon tests of the samples. The tensile strength of the tested rebars was 437.5 MPa. Three specimens were also tested to determine
the concrete compressive strength; an average compressive strength of 33.1 MPa was measured. Figure 9 and Table 1 describe the material
properties of composite beams that were instrumented for loading with an actuator of 2,000 kN capacity located 1.78 m from the foundation
surface. The length of the beam and height of the column component were 2 m and 1 m, respectively, as shown in Figure 9. Description of the
three test specimens is shown in Table 1. Column components of Specimens 1 and 2 were assembled using sleeves (Figure 9 (b)), and columns were
fabricated by pressure welding (Figure 9 (c)) for Specimen 3. The beam‐column joint was designed based on simplified method based on strain
compatibility described in Section 6.2. The post‐yield behavior was, then, validated by nonlinear FEA in Sections 6.5 and 6.6.

4.2 | Instrumentation and test set‐up


The test set‐up of the specimens is illustrated in Figure 10, and the cyclic loading protocol is depicted in Figure 11; two or three cycles were
applied for each stroke length consisting of virgin and stabilized cycles. The applied lateral displacements were estimated by calculating the
product between the interstory drift angle (rad) and the height of the specimen measured from the slab surface. Figure 11 illustrates the story
drift ratios (%) with the associated lateral displacements and the number of cycles applied for each displacement (AISC[31]). As many as 80 strain
gauges were installed on each test specimen; the gauges were attached to the top/bottom/web of flanges and to the reinforcement. The
measured strains, which were compared to the nonlinear, inelastic, FEA, were used to explore the influence of the unsymmetrical steel on
the concrete section.

4.3 | Experimental investigation


Figures 12 and 16 display the major failure modes of the composite beams. The flexural moment capacity of the beam section up to 182.5 kN‐m in
the direction along which the major concrete was unbonded was measured with a steel flange in the compression zone; no steel flange was
available in compression. However, 208.3 kN‐m was reached when the major concrete was bonded with a steel flange in the compression zone;
the steel flange was available in compression. The precast composite beams with T‐section steel showed a significant increase in flexural moment
capacity with good ductility in addition to providing steel joint connection to the columns, which would enhance the erection capacity to match
that of steel frames.

5 | A N UM E R I C A L M O DE L

5.1 | 3D mesh for FEA


Table 2 presents the material properties used in FEA and the test specimen. The 3D meshes used for the FEA are shown in Figure 13. The section
of the steel beam shown in Figures 9 and 14 is not symmetrical with the loading, yielding different flexural capacity and damage characteristics in
each direction.

FIGURE 8 Fabrication of specimens


NZABONIMPA ET AL. 9 of 25

(a) Section properties

(b) Column connection by sleeve

FIGURE 9 Section properties and geometry


of the specimens used for experimental
investigation (Hong et al.[22]). (a) Section
properties, (b) column connection by sleeve,
and (c) column connection by pressure
welding (c) Column connection by pressure welding

5.2 | Parameters for nonlinear numerical model


The stiffness degradations, strains of rebar, cracking/crushing of the concrete, and yielding/fracture of the steel beam sections based on the
interaction between concrete and steel were implemented in the evaluation of the seismic performance of unsymmetrical steel–concrete
composite precast beams with T‐section steel, replacing conventional cast‐in‐place composite beams. Parameters defining plastic flow potential
function and concrete‐damaged plasticity in the nonlinear FE model are presented in Table 3, including dilation angles, viscosity, and eccentricities
that were used to estimate the seismic capacity of the precast composite sections. The constitutive relationship between concrete and damage
variable as a function of plastic strain is suggested in Figure 15 (a–d), by which the degradation of concrete both in compression and tension
are governed.
10 of 25 NZABONIMPA ET AL.

FIGURE 10 Test set‐up and loading application of the specimen

FIGURE 11 The cyclic loading protocol


(AISC[31])

FIGURE 12 Failure modes of specimens

Concrete under compression was modeled for all specimens using the Kent‐Park constitutive equation. The first part of the curve is the
linear‐elastic region, which ends at the stress level of σco = 0.4f′c,as shown in Figure 15 (a). The compression damage parameter for concrete under
uniaxial compression loading, which was implemented in FEA, is introduced to model concrete‐damaged plasticity based on the stiffness
NZABONIMPA ET AL. 11 of 25

TABLE 2 Material property used in finite element analysis


Size Material properties

Concrete beam 300 × 350 (mm2) f′c = 33.1 MPa


2
Concrete column 550 × 550 (mm ) f′c = 33.1 MPa
T‐steel (SM490) 150 × 150 × 6.5 × 9 (mm3) Fy = 325 MPa; εy = 0.0016
Bolt M22‐F10 T Fu = 1000 MPa; εy = 0.0045
Stud bolt D13 (height: 53 mm) fy = 400 MPa; εy = 0.0019
Rebar D22, D13 fy = 400 MPa; εy = 0.0019

FIGURE 13 3D model of the test specimen and 3D mesh of the test specimen

FIGURE 14 Headed studs installation

degradation variable shown in Figure 15 (b). This commences at the peak of the constitutive equation of concrete. The tensile constitutive
relationships for concrete and the damaged plasticity that defines the softening part of the tensile constitutive relationship are also shown in
Figure 15 (c,d). The post‐yield behavior based on von Mises yield criterion for steel section was compared with that obtained using Tresca model
for the specimen #1 in Figure 18. Von Mises criterion is less conservative than Tresca, overestimating yield criterion because von Mises model
predicts lower effective stress (lower actual comparison to the yield strength) than Tresca for a given status of stress. Load displacement
relationship computed with von Mises yield criteria was slightly lower than that based on Tresca criteria. The discrepancies between numerical
analysis and test data could also be justified by the monotonic load application through the FE study, while test was performed based on cyclic
loads. The cyclic degradation of the materials occurred during testing were not fully reflected in the numerical analysis. The implemented FE model
showed a good correlation with cyclic test data up to the maximum loads despite the cyclic degradation of the materials and mechanical properties.

6 | E V A L U A T I O N O F S E I S M I C P E R FO R M A N C E

6.1 | FE study of seismic behavior of unsymmetrical precast composite beams


The deflected shape of the steel–concrete composite beams based on the damage‐plastic model of concrete is demonstrated in Figure 16. The
structural characteristics differ based on direction, because the specimens are not symmetrical. The evaluated seismic response of the specimens
12 of 25 NZABONIMPA ET AL.

TABLE 3 Parameters identified for finite element analysis of an unsymmetrical composite beam
Dilation angle Damage variables (%) Eccentricity fbo/fco Kc Viscosity parameter

First quadrant
30 (Legend 6) D100 0.1 1.16 2/3 0.001
35 (Legend 7) D100 0.1 1.16 2/3 0.001
30 (Legend 8) D100 0.1 1.16 2/3 0.004
30 (Legend 9) D100 0.1 1.16 2/3 0.007
Third quadrant
35 (Legend 4) D10 0.1 1.16 2/3 0.001
40 (Legend 5) D10 0.1 1.16 2/3 0.001

FIGURE 15 The constitutive relationship of concrete and damage variables; (a) constitutive relation of concrete for compression, (b) damage
plasticity of concrete for compression, (c) damaged plasticity of concrete for tension, and (d) damaged plasticity of concrete for tension

is presented in Figure 17, where a good correlation between numerical estimation and test observation was found. The strains of concrete, rebars,
and steel section are presented on the load displacement curve. It is noted that a lateral load capacity of 83.4 kN and 96.0 kN for load reversals
corresponding to a concrete strain of 0.003 were identified in Figure 17. The numerical estimation without damaged plasticity demonstrates
significant inconsistency with observed performance of the specimens. The parameters for a steel–concrete precast composite beam consisting
of T‐section steel are listed in Table 3. The magnitude of the damaged plasticity variable should be calibrated with care for the estimation of the
flexural capacity of steel–concrete composite beams, especially when the composite beams are fabricated with unsymmetrical steel sections.
The selection of a constitutive relationship for concrete based on a damage variable with appropriate magnitude should reflect how the concrete
is bonded with the steel sections in the compression zone because the nonlinear parameters are influenced by the ductility of the composite
sections. Results from the analytical estimation of test specimens represented by the curve of Legend 6 in the first quadrant (Figure 17), where
no steel flange was available for bonding to the major concrete section in the compression zone, are in agreement with test data when the full
damaged plasticity introduced in Figure 15 was used with a dilation angle of 30o and viscosity of 0.001. It is demonstrated that the flexural capacity
with a dilation angle of 35o represented by Legend 7 in Figure 17 is greater by 3.7% than the computed capacity with a dilation angle of 30o, as
shown in Figure 17. However, the curve represented by Legend 6 is small relative to the test data in the third quadrant, where a steel flange
was available for bonding with the major concrete section in the compression zone as shown in Figure 17. The curve represented by Legend 4
of the load–displacement relationship is in agreement with the test data in the third quadrant, which displays the composite action with the
concrete, allowing the compressive steel flange strains to reach 0.0014 and 0.0029 for compressive concrete strains of 0.0056 and 0.008, respec-
tively, in all specimens shown in Figures 17--20. This result indicates that a smaller damage variable and large dilation angle are preferable for
composite beam sections where major concrete sections are bonded with steel flanges; however, the numerical analysis of the curve of Legend
NZABONIMPA ET AL. 13 of 25

FIGURE 16 Unsymmetrical deflected shapes for the tested specimen. (a) Load applied toward left steel (no flange available in compression) and (b)
load applied toward right (steel flange available in compression)

FIGURE 17 Sensitivity of dilation angles and damage parameter to the flexural capacity of the composite beam; Legends 4–7

4 yields values greater than the experimentally observed data with a large discrepancy in the first quadrant when the 10% damaged plasticity model
with a dilation angle of 35o is implemented in Curve 4, as shown in Figure 17. The estimate and observed strains are compared in Figure 21,
demonstrating that the analytical data based on the FE and strain compatibility are in agreement with the test data. Numerical analysis can serve
as an effective alternative for the design of composite beams with high accuracy.

6.2 | Simplified method based on strain compatibility


The load–strain (rebars) relationship represented by Legend 10 of Figure 21 was obtained by simplified method based on strain compatibility, which
allows for the precise prediction of the nominal moments at all limit states including yield and maximum load limit state; the flexural moment
capacities were calculated based on the neutral axis determined from the equilibrium equations at each limit state.
From the equilibrium condition when loads were applied with negative direction, the location of the neutral axis (c) from the top of the section
can be arranged at yield limit state as Equation 10. Solving the quadratic equation (10) for c, the location of the neutral axis from the top of the
beam was obtained as 122 mm at yield limit state. The corresponding nominal moment capacity (Equation 11) of the composite beam was
calculated to be 182 kN‐m. The concrete strain at yield limit state, εc = 0.00160, corresponding to εt = 0.00219 was calculated through the iteration.
14 of 25 NZABONIMPA ET AL.

FIGURE 18 Test data versus nonlinear numerical results (Specimen PS1); Legends 4–7

FIGURE 19 Test data versus nonlinear numerical results (Specimen PS2); Legends 4–7

The nominal moment capacity of the composite beam was calculated as 198.8 kN‐m at the maximum load limit state.

( )
1 1 ðεsy Þ2
αf ′c b þ tw εc Es þ tw Es þ tw Fy c2
2 2 εc

n    o
þ A′s εc Es þ A′f εc Es −tw εc Es d′′′ þ tf −As f y −tw F y d−d′ c

   2
tf 1
þ −A′s εc Es d′′ −A′f εc Es d′′′ þ þ tw εc Es d′′′ þ tf ¼ 0; C ¼ 122:0 mm; (10)
2 2
NZABONIMPA ET AL. 15 of 25

FIGURE 20 Test data versus nonlinear numerical results (Specimen PS3); Legends 4–7

εc  2 εc   tf

tf

tf

c−d′′ þ A′f Es
Mn ¼ αf ′c bcðc−γcÞ þ A′s Es c−d′′′ −tf c−d′′′ − þ c−d′′′ −
c c 2 2 3
!
1 ′ εc  ′′′
2 1 ′ εsy 1 ðεsy Þ2 (11)
þ Aw Es c−d −tf þ As f y ðd−cÞ þ Awp F y d−c−d þ c þ Awny Es c
3 c 2 εc 3 εc
¼ 182:4 kN−m;

   
εsy εsy
where, A′w ¼ tw c−d′′′ −tf ; Awp ¼ tw d−c−d′ − ; Awnp ¼ tw c
εc εc
Figure 21 (b,c) shows the strain and the stress of the composite beam section at the yield and maximum load limit state, showing the neutral
axis, corresponding stress and equilibrium of internal forces of composite sections. The simplified method based on strain compatibility is in
agreement with the full scale test data (Curve 1) and result obtained by nonlinear FEA (Curves 4 for the third quadrant and 6 for the first quadrant)
as shown in Figure 21 (a).

6.3 | Sensitivity of dilation angles and damage variable to the nonlinear numerical behavior of the composite
beams
The structural behavior of the specimens was accurately estimated by the proposed model using a dilation angle of 30o and full damage
application in the region where no steel sections were available for bonding with the concrete section in compression. As a result, less ductility
was demonstrated than in the region where the concrete was bonded with the steel flange in compression. Large concrete‐damaged plasticity
and smaller dilation angles are recommended for the analytical investigation of steel–concrete composite beams when the concrete is less well
bonded with the steel sections in compression. The load–displacement relationship of the specimens correlates well with small damage
parameters when steel sections encased in concrete contributed to the prevention of concrete damage, indicating that greater dilation angles
and lower damage variables are optimal for regions in which concrete is bonded strongly with steel sections in compression, thus increasing
concrete ductility.

6.4 | Viscosity
The use of viscosity will enhance the convergence of the solution by reducing the plastic flow rate and plastic degradation, leading to a
decreased rate of plastic fracture within the section. Table 4 compares the sensitivity of the viscosity to the execution time. Note that the
computing time is enhanced and the strokes are calculated further. The computing speed is substantially enhanced from 1.43 mm/h to
4.8 mm/h when a viscosity of 0.007 was implemented instead of 0.001, as shown in Figure 22 and Table 4. As shown in Table 4, the analysis
results also increase slightly by 8.6% with significant increases in viscosity. However, some studies state that changes in viscosity do not
compromise the analysis results.
16 of 25 NZABONIMPA ET AL.

(a) Legends 1, 4, 5, 6, 7, 10 (compatibility)

(b) At yield limit state

FIGURE 21 Test data versus numerical data


(based on finite element analysis and strain
compatibility). (a) Legends 1–7, 10
(compatibility), (b) at yield limit state, and (c) at
(c) At maximum load limit state maximum load limit state

TABLE 4 Run efficiency versus viscosity parameter

ABAQUS model Viscosity parameter CPU/run efficiency/


Legend 6 0.001 62.2 hr (2.6 days)/1.43 (mm/h)
(reached stroke: 89.0 mm, reached force: 104.2 kN)
Legend 8 0.004 26.9 hr (1.1 days)/3.3 (mm/h)
(reached stroke: 89.0 mm, reached force: 109.2 kN; force increased by 4.8%)
Legend 9 0.007 22.5 hr (0.9 days)/4.8 (mm/h)
(reached stroke: 106.9 mm, reached force: 113.2 kN; force increased by 8.6%)

6.5 | Influence of steel section on the activation of rebars and concrete in compression for damage assessment
The compressive strains of concrete and rebars increased similarly up to a concrete strain of 0.003 in both directions of the loading application
(Legends 4 and 6 of Figure 17); however, it is demonstrated that the strain rates of concrete and rebar in the compression zone were increasing
NZABONIMPA ET AL. 17 of 25

FIGURE 22 Influence of viscosity on the flexural capacity of the unsymmetrical composite beams; Legends 1 (PS2, Figure 19), 6, 8, 9

FIGURE 23 Influence of steel flange on the strain rates of concrete and rebars (compression); (a) concrete, (b) Rebars, (c) steel web, and (d) steel
flange

more significantly in the third quadrant (Legend 4 of Figure 17) where a steel flange was integrated with the concrete than in the first quadrant
(Legend 6 of Figure 17) where the concrete was not bonded with steel sections after undergoing strains of 0.006 and 0.002 for the concrete
and rebars, respectively, as shown in Figure 23 (a,b). The compressive concrete strain remained constant in the region where the concrete was
not bonded with steel sections, while those in the region where the steel sections were strongly bonded in compression were constantly increasing;
this is reflected in the numerical parameters, which are influenced by increased concrete ductility with greater dilation angle and less damage
variable. The compressive strain rates of the steel flange shown in Figure 23 (d) increased linearly from the stroke of 18 mm and 94.4 mm; herein,
larger concrete strains were sustained as demonstrated in Figure 23 (a). However, the compressive strain rates of the steel web in the first quadrant
were larger where the steel flange was not available, as shown Figure 23 (c).
Figure 24 (a) shows rebars were activated more than steel web section; the use of rebars with steel section contributed to the flexural capacity
effectively, maximizing the interaction of the steel section and reinforced concrete when first quadrant is in compression (Legend 6 of Figure 17).
18 of 25 NZABONIMPA ET AL.

FIGURE 24 Strain–stroke of the selected structural components; activation of structural members (a) when first quadrant is in compression;
compressive strain (Legend 6 of Figure 17), (b) when third quadrant is in compression; compressive strain (Legend 4 of Figure 17), (c) when first
quadrant is in compression; tensile strain (Legend 6 of Figure 17), and (d) when third quadrant is in compression; tensile strain (Legend 4 of Figure
17)

Figure 24 (b) plots strain–stroke relationship in which similar rates of compressive strain increase were found for steel flange, headed studs, and
steel web when third quadrant is in compression (Legend 4 of Figure 17). The rates of strain increased rapidly after the strain corresponding to
concrete strain of 0.003. Rebars were activated quickly, providing contribution to the flexural capacity effectively while the used of steel sections
were saved. Rates of tensile strain increase were found similar for steel flange, headed studs, and steel web when both first and third quadrants are
in compression (Legends 4 and 6 of Figure 17), as shown in Figure 24 (c,d).
Figure 25 presents stress and strain evolution in the hybrid composite beam at the concrete compressive strains of 0.003 and 0.008; numerical
evaluation of the seismic performance of the proposed composite structure is in good agreement with the test data. When the concrete
compressive strain reached 0.003, the tensile strains of the rebar and steel flange were 0.002 and 0.0015 (Figure 25 (a); 1), respectively, when first
quadrant is in compression (Legend 6 of Figure 17), while the tensile strains of the rebar and steel web were 0.0018 and 0.0011 (Figure 25 (b); (1)),
respectively, when third quadrant is in compression (Legend 4 of Figure 17), as shown in Figures 17 and 25. These numerical strains were well
compared with test data shown in Figure 21, where the experimental load–strain relationship was retrieved. In Figure 17, it was found that the
compressive concrete strains increased to 0.006 and 0.0061 in the first quadrant, respectively, and to 0.0056 and 0.008 in the third quadrant when
rebar tensile strain increased to 2εy and 5εy, respectively. The greater tensile strain in the first and greater compressive strain in the third quadrant
contributed to a greater flexural capacity along with the direction causing the steel flange to be in compression. It is shown that concrete strains
greater than those allowed in most codes can be utilized when the concrete sections are bonded with a steel section, as in the steel–concrete
composite members. The effective use of the two materials, steel and concrete, can provide greater seismic capacity with greater ductility,
contributing to a maximized structural capacity within the composite precast beams.
Figure 26 illustrates test results at the drift angle of 0.03o with cracks observed along the height of the beams (3–4 mm wide cracks in the
bottom of specimens and 0.1–0.2 mm wide cracks at the top). Cracks were more dominant when loads were applied with the negative direction
along which the concrete in tension zone was not confined by the steel sections, whereas the major concrete section in the compression zone
was confined by steel flange. However, more composite action with the concrete confined by steel flange in compression zone allowed concrete
strains to reach 0.0056 and 0.008. The strains of compressive steel flange also reached 0.0014 and 0.0029, contributing to the greater flexural
NZABONIMPA ET AL. 19 of 25

(a) When first quadrant is in compression (Legend 6 of Figure 17)

(b) When third quadrant is in compression (Legend 4 of Figure 17)

FIGURE 25 Stress evolution in the hybrid composite beam at concrete compressive strains of 0.003 and 0.008; (a) when first quadrant is in
compression (Legend 6 of Figure 17) and (b) when third quadrant is in compression (Legend 4 of Figure 17)

strength of the composite section with the loads of negative direction. It is worth noting that more hysteretic energy was also dissipated when the
concrete was confined by steel flange in compression zone as shown in the observed test cycles of Figure 27 and Table 5.

6.6 | Prediction of crack propagation for tensile cracking and compressive crushing; valuation of damage
evolution
The failure modes with cracks and spalled concrete were analytically estimated based on the concrete plasticity defined in Sections 2.2 and
5.2, in which the estimation of crack propagation is in agreement with photos taken during testing, as shown in Figure 28. Strains and failure
modes were also accurately identified at the concrete, rebars, and steel sections of the steel–concrete composite beams in Figure 27, demon-
strating the influence of the interactive behavior between steel and concrete materials on the flexural capacity of precast composite beams. In
the interest of simplicity, the term “crack” is used to indicate a direction in which the maximum principal tensile stress exceeds the tensile
strength of the brittle material as defined in the constitutive equation. The directions of the crack surface normal to the existing cracks close
and reopen. Rebar elements based on the one‐dimensional strain theory are defined as embedded in oriented surfaces. The rebars are
assumed to be independent of the concrete cracking, because they are typically used with elastic‐plastic material behavior and are superposed
20 of 25 NZABONIMPA ET AL.

FIGURE 26 Failure mode at drift angle of θ: 0.03°; (a)−1 failure mode (positive direction) at drift angle of θ: 0.03°, (a)−2 failure mode (negative
direction) at drift angle of θ: 0.03°, (a) specimen #1, (b)−1 failure mode (positive direction) at drift angle of θ: 0.03°, (b)−2 failure mode (negative
direction) at drift angle of θ: 0.03°, (b) specimen #2, (c)−1 failure mode (positive direction) at drift angle of θ: 0.03°, (c)−2 failure mode (negative
direction) at drift angle of θ: 0.03°, and (c) specimen #3

on a mesh of plain concrete. In Figure 28, concrete cracks for compression in both quadrants were calculated based on the concrete‐damaged
plasticity model, which has the potential to represent completely nonlinear, inelastic failure mechanisms of concrete with tensile cracking and
compressive crushing. The concrete spalling was evaluated via numerical investigation in both compression and tension; the results were
validated via experimental observations.
NZABONIMPA ET AL. 21 of 25

FIGURE 26 Continued.

7 | DISCUSSIONS

This study proposed steel–concrete composite beams featured efficient interaction between the two materials for a good seismic performance.
The use of integrated composite structures is beneficial for use in buildings when replacing conventional cast‐in‐place composite beams. The
capability of reduced floor height by locating T‐section steel inside the slab concrete was also demonstrated. However, the access by engineers
to a sophisticated analysis and the use of complex nonlinear FEA parameters for the analysis and design of steel–concrete composite members
22 of 25 NZABONIMPA ET AL.

(a) Damage evolution

(b) Damage description

FIGURE 27 Concrete damage in compression with identified strains; (a) damage evolution and (b) damage description

TABLE 5 Summary of test and numerical data for all specimens


Yield limit state Maximum load limit state
Specimens Test Calculated Error (%) Test Calculated Error (%)

Specimen #1 (PS1) Positive moment Load (kN) 80.4 83.1 −3.2 98.3 95.2 +3.3
Rebar strain 0.00216 0.00219 −1.4 0.00786 0.00786 0
Negative moment Load (kN) 115.2 111.5 +3.3 124.3 120.4 +3.2
Rebar strain 0.00220 0.00219 +0.5 0.00517 0.00482 +7.2
Specimen #2 (PS2) Positive moment Load (kN) 83.9 83.1 +1.0 103.2 95.2 +8.4
Rebar strain 0.00219 0.00219 0 0.00898 0.00786 +14.2
Negative moment Load (kN) 112.3 111.5 +0.7 121.7 120.4 +1.1
Rebar strain 0.00217 0.00219 −0.9 0.00500 0.00482 +3.7
Specimen #3 (PS3) Positive moment Load (kN) 83.3 83.1 +0.2 98.0 95.2 +2.9
Rebar strain 0.00214 0.00219 −2.3 0.00760 0.00786 −3.3
Negative moment Load (kN) 109.7 111.5 −1.6 126.4 120.4 +5.0
Rebar strain 0.00216 0.00219 −1.4 0.00490 0.00482 +1.7
Average Positive moment Load (kN) 82.5 83.1 −0.7 99.8 95.2 +4.9
Rebar strain 0.00216 0.00219 −1.2 0.00815 0.00786 +3.6
Negative moment Load (kN) 112.4 111.5 0.8 124.1 120.4 +3.1
Rebar strain 0.00218 0.00219 −0.5 0.00502 0.00482 +4.2
NZABONIMPA ET AL. 23 of 25

FIGURE 28 Degradation of the steel–concrete composite beam based on damaged plasticity

are limited, which made the design codes conservative. This study suggested combinative parameters of the dilation angles, damaged variables, and
viscosity to find accurate and fast convergence for the design of the steel beams encased by structural concrete. This study selected a dilation angle
of 35o combined with a 10% damage variable based on the failure modes observed for concrete strongly bonded with steel flanges (both flange and
web). The combination of a dilation angle of 30o and a 100% damage variable was implemented, reflecting more brittle failures of concrete that was
less well bonded with steel sections (web only). The simplified design method for steel–concrete composite members based on strain compatibility
was also presented.
For the unsymmetrical composite beams introduced in this study, the numerical evaluation of brittle concrete that is less bonded with steel
sections (web only) must be based on the full stiffness degradation variable. Lower stiffness degradation for the concrete with less damage must
be implemented in concrete sections that are strongly bonded with steel sections (both flange and web) to achieve an accurate analytical evaluation
of structural performance of steel–concrete composite beams. In the compression zone, the composite action of steel flanges with concrete was
demonstrated, in which the compressive steel flange strains were 0.0014 and 0.0029 corresponding to the compressive concrete strains of 0.0056
and 0.008, respectively, as shown in the third quadrant of Figure 17. This problem arises with unsymmetrical beams sections. Both the predicted
flexural capacity and the crack patterns are in agreement with the experiment results obtained by the authors. The influence of the viscosity
parameter and rate of plastic strains on the solution accuracy and convergence for FE solutions was also studied, and it was determined that
the rate of convergence and computation efficiency were significantly improved as viscosity increased. The increase in viscosity can lead to over
estimation of the structural capacities of the proposed hybrid steel–concrete beams, as shown in curves represented by Legends 9 and 6 in Table 3
and Figure 22. This indicates that viscosity should be carefully selected, because large viscosity values can influence FE results. Good agreement
with test data was obtained when a value of 0.001 was used for the viscosity parameter, as illustrated by the curve in Legend 6 in the first quadrant
of Figure 17.

8 | C O N CL U S I O N

The seismic performance of steel–concrete composite precast beams with T‐section steel based on the concrete‐damaged plasticity was evaluated,
expecting to contribute to the fast and effortless construction of the precast composite steel frames. Many construction sites of composite frames
depending on the conventional method and cast‐in‐place frames were replaced by the precast composite frames proposed in this study. Numerical
investigation on steel frames encased by precast structural concrete was also established for an application of the precast composite frames to
replace the conventional buildings. For this to be realized, FE model with damage parameters for the evaluation of the nonlinear, post‐yield
behavior of the precast composite steel beams should be available. This study provided practical design parameters and design recommendations
by examining the influence of precast composite beams with unsymmetrical steel sections on the concrete degradations and damage evolution.
Evaluation of the seismic capacity of steel–concrete composite precast beams with T‐section steel was performed via numerical investigation
and verified by the earlier experimental study of the authors. This structural solution combines steel and concrete, offering advantages not only
in terms of structural stability but also constructability, cost‐effectiveness, and environmental friendliness. The proposed hybrid composite struc-
tural system maintains the advantages of both precast concrete structures and steel frames; herein, the composite precast concrete frames are cre-
ated with steel joints, combining the economy of concrete construction with the speed of steel construction. The influence of the stiffness
degradation variable and dilation angle on the nonlinear plastic deformation of steel–concrete precast composite beams was studied to evaluate
their nonlinear structural behaviors. Parameters including stiffness degradation variables and dilation angles were also calibrated to provide an
24 of 25 NZABONIMPA ET AL.

accurate estimation of the structural behavior of steel–concrete precast composite beams. This study summarized the FEA‐based material models
with the identified parameters. Ductility with confining pressures defining dilation angles and stiffness degradation was an important parameter
when numerically determining the flexural capacity of the composite beams. Specified nonlinear parameters accounting for ductility and degree
of composite between steel and concrete are recommended for the structural geometry and confining pressures. The evaluation of the seismic
capacity of unsymmetrical steel–concrete composite precast beams was performed both experimentally and numerically. This study is also dedi-
cated to the elimination of high‐cost experiments via nonlinear numerical technique with the proposed parameters, offering low‐cost but accurate
numerical investigation for the design of complicated composite structures that are not limited to the unsymmetrical steel–concrete precast com-
posite beams introduced in this study.

ACKNOWLEDGEMEN TS
This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the
Ministry of Education (2016R1D1A1A02937558).

ORCID

Won‐Kee Hong http://orcid.org/0000-0002-1955-9598

RE FE R ENC E S
[1] A. Prakash, N. Anandavalli, C. K. Madheswaran, J. Rajasankar, N. Lakshmanan, Int. J. Mec. Appl. 2011, 1(1), 1.
[2] Wahalathantri BL, Thambiratnam DP, Chan TH, Fawzia S (2011). A material model for flexural crack simulation in reinforced concrete elements using
ABAQUS. In Proceedings of the First International Conference on Engineering, Designing and Developing the Built Environment for Sustainable
Wellbeing, Queensland University of Technology, Australia. 2011, 260–264.
[3] P. Grassl, M. Jirásek, Int. J. Sol. Struct. 2006, 43(22), 7166.
[4] M. Szczecina, A. Winnicki. Calibration of the CDP model parameters in ABAQUS. The 2015 International Conference on Advances in Structural
Engineering and Mechanics, Computational Technologies in Concrete Structures, Incheon, Republic of Korea. 2015.
[5] S. A. Kulkarni, B. Li, W. K. Yip, J. Constr. Steel Res. 2008, 64(2), 190.
[6] H. H. Korkmaz, T. Tankut, Eng. Struct. 2005, 27(9), 1392.
[7] A. S. Genikomsou, M. A. Polak, Eng. Struct. 2015, 98, 38.
[8] O. Omidi, V. Lotfi, Int. J. Civil Eng. 2010, 8(3), 187.
[9] J. Lee, G. L. Fenves, J. Eng. Mech. 1998, 124(8), 892.
[10] R. A. Hawileh, A. Rahman, H. Tabatabai, App. Math. Model. 2010, 34(9), 2562.
[11] R. Malm, Predicting shear type crack initiation and growth in concrete with non‐linear finite element method, Doctoral thesis, School of Architecture
and the Built Environment, Sweden, 2009.
[12] W. B. Bickford, Mechanics of Solids: Concepts and Applications, Irwin, Boston 1997.
[13] C. Ansel, S. Ugural, K. Fenster, Advanced Mechanics of Materials and Applied Elasticity, Pearson Education, Inc, Boston, US 2011.
[14] C. Amadio, C. Bedon, M. Fasan, M. R. Pecce, Eng. Struct. 2017, 138, 394.
[15] J. M. Ricles, S. D. Paboojian, J. Structur. Eng. 1994, 120(8), 2474.
[16] R. Kindmann, R. Bergmann, L. G. Cajot, J. B. Schleich, J. Constr. Steel Res. 1993, 27(1–3), 107.
[17] S. Ahmad, A. Masri, Z. A. Saleh, Alex. Eng. J. 2017.
[18] H. Hou, X. Liu, B. Qu, T. Ma, H. Liu, M. Feng, B. Zhang, Eng. Struct. 2016, 126, 405.
[19] J. D. Nzabonimpa, W. K. Hong, S. C. Park, Struct. Design Tall Spec. Build. 2017, 26(1).
[20] C. Dundar, S. Tokgoz, A. K. Tanrikulu, Build. Environ. 2008, 43(6), 1109.
[21] W.‐K. Hong, S.‐c. Park, H.‐C. Lee, J.‐M. Kim, S.‐I. Kim, S.‐G. Lee, K.‐J. Yoon, Struct. Design Tall Spec. Build. 2009, 19(6), 679.
[22] W. K. Hong, S. Y. Jeong, S. C. Park, J. T. Kim, Energ. Buildings 2012, 46, 37.
[23] J. Lubliner, J. Oliver, S. Oller, E., Int. J. Sol. Struct.. 1989, 25(3), 299–326.
[24] J.C. Galvez, J. Cervenka, D.A. Cendon, V. Saouma (2002) Cem. Concr. Res. 2002, 32 (10), 1567–1585.
[25] W.‐F. Chen, D. J. Han, Plasticity for structural engineers, Gau Lih Book Co., Ltd, Taipei, Taiwan 1995.
[26] Hibbit, Karlsson & Sorensen, Inc., Pawtucket, R. I, ABAQUS User's manual version 6.7, 2007.
[27] R. Malm (2006), Shear cracks in concrete structures subjected to in‐plane stress, Licentiate thesis, School of Architecture and the Built Environment,
Sweden, 2006.
[28] Dassault Systèmes Simulia Corp., ABAQUS analysis User's manual 6.14‐2, Providence, RI, USA, 2014.
[29] S. V. Chaudhari, M. A. Chakrabarti, Int. J. Comp. Appl. 2012, 44(7), 14.
[30] R. Park, T. Paulay, Reinforced concrete structures, John Wiley & Sons, Inc 1975.
[31] AISC/ANSI 358, Prequalified connections for special and intermediate steel moment frames for seismic applications specification. Chicago (IL):
American Institute of Steel Construction, Inc, 2005.
NZABONIMPA ET AL. 25 of 25

Mr. J.D. Nzabonimpa is currently enrolled as a Ph.D. candidate in the Department of Architectural Engineering at Kyung Hee University,
Republic of Korea. His research interest includes precast composites structures.

Dr. Won‐Kee Hong is a professor of Architectural Engineering at Kyung Hee University. Dr. Hong received his Master's and Ph.D. degrees
from UCLA, and he worked for Englelkirk and Hart, Inc. (USA), Nihhon Sekkei (Japan), and Samsung Engineering and Construction Company
(Korea) before joining Kyung Hee University (Korea). He also has a professional engineering license from both Korea and the USA. Dr. Hong
has more than 30 years of professional experience in structural engineering. His research interests include a new approach to construction
technologies based on value engineering with hybrid composite structures. He provided many useful solutions to issues in current structural
design and construction technologies as a result of his research that combines structural engineering with construction technologies. He is
the author of numerous papers and patents both in Korea and the USA. Currently, Dr. Hong is developing new connections that can be used
with various types of frames including hybrid steel–concrete precast composite frames (SMART frames), precast frames, and steel frames.
These connections would contribute to the modular construction of heavy plant structures and buildings as well.

Mr. Jisoon Kim received his Master's degree from the Department of Architectural Engineering at Kyung Hee University, Republic of Korea.
His research interest includes precast composites structures.

How to cite this article: Nzabonimpa JD, Hong W‐K, Kim J. Strength and post‐yield behavior of T‐section steel encased by structural
concrete. Struct Design Tall Spec Build. 2018;27:e1447. https://doi.org/10.1002/tal.1447

You might also like