Numerical and Analytical Methods For The Design of Beam-Column Sub-Assemblies With Composite Deck Slab - 2022
Numerical and Analytical Methods For The Design of Beam-Column Sub-Assemblies With Composite Deck Slab - 2022
https://doi.org/10.1007/s10518-022-01500-1
ORIGINAL ARTICLE
Abstract
A finite element numerical model is used to estimate the force displacement backbone
curve of steel moment frame beam-column-joint subassemblies tested at large scale under
repeated cyclic loading with different composite deck slab configurations. The slab con-
figurations included: no-slab; fully isolated slab; two configurations for detailing the slab
near the region between the column flanges; and a full depth reinforced slab. Then, a sim-
ple analytical model was developed so that designers can estimate the likely peak strength
due to slab effect. This analytical model considers the common deformation modes and
the strength hierarchy. It was found that the finite element numerical model captured the
backbone envelope of experimental tests done on different slab configurations for the bare
frame, isolated slab and full depth slab configurations, but overestimated the strength at
larger displacements for the other configurations due to difficulty in considering the slab
modes of failure. The simple analytical model considered the nonlinear deformation modes
of steel beam plastic hinging, concrete crushing both outside the column flange and within
the beam flanges, slab shear fracture between the column flange tips, slab longitudinal and
lateral reinforcement yielding as it carried the subassembly moments and transferred forces
between the steel frame and the slab, and shear stud deformation. The proposed analytical
model matched the experimental strengths and failure modes. The proposed finite element
model is suitable for research, and the analytical model matched the experimental results,
and is suitable for consideration in design.
Keywords Force transfer mechanism · Micro model · Moment frames · Shear studs ·
Structural steel · Composite slabs
1 Introduction
Steel framed structures with composite slabs are popular in many seismically active
regions around the world. Following the 2010–2011 Canterbury earthquakes, the num-
ber of steel-framed buildings with composite deck slabs in New Zealand has dramati-
cally increased because of their: proven seismic performance; architectural flexibility;
* Tushar Chaudhari
Extended author information available on the last page of the article
13
Vol.:(0123456789)
Bulletin of Earthquake Engineering
13
Bulletin of Earthquake Engineering
detailed so that it does not degrade in strength during seismic motions and a smaller steel
beam may be used.
Based on the above discussion, there is a need to better understand the slab interaction
effects in steel moment-frames with conventional slabs, isolated slab, or full depth slab
considerations, and to be able to analyse their performance so that appropriate design deci-
sions can be made. This paper seeks to address this need by seeking answers to the follow-
ing questions:
(i) Does numerical simulation, using finite element analysis of the beam-column subas-
sembly, capture the force transfer mechanisms associated with slab-column interac-
tion?
(ii) Can simple hand calculation methods be used to estimate composite beam-column
slab subassemblies lateral strength and stiffness?
(iii) What modes of deformation control subassembly strength?
(iv) What are the implications of the findings of this work for design?
Figure 1 explains two force transfer mechanisms between the slab and joint defined by
Eurocode EN1998-1 (2004), as well as possible modes of failure within the slab-column
interaction zone. The first lateral force resisting mechanism, Mechanism 1, involves direct
compression from the slab on the outside of the column flanges, while Mechanism 2
involves compression on the inside of the column flange and inclined concrete compression
struts. A third mechanism, Mechanism 3, also exists as a result of out-of-plane beam shear
studs causing the top of the beam to move, thereby inducing beam torsion, and in-plane
joint moment. While in many cases the contribution of force transfer from this mechanism
is small, (Webb et al. 2018) it is discussed further here.
13
Bulletin of Earthquake Engineering
Load paths for the slab to contribute to the frame sub-assembly lateral strength are as
follows:
a. Load path 1 (LP1, involving Mechanism 1 as shown in Fig. 2a): Force is carried by
the shear studs on the left-hand side of the column (i.e. sagging side) (mode 1) as they
carry force from the slab to the beam (mode 5), and by the concrete at the column outer
flange-slab interface (mode 2),
b. Load path 2 (LP2, involving Mechanism 1 as shown in Fig. 2b): Force is carried by
mode-2, mode 5, and tie reinforcement tension (modes 6 and 7) and then through the
shear studs on the right-hand side of the column (i.e. hogging side) (mode 1).
c. Load path 3 (LP3, involving Mechanism 2 as shown in Fig. 2c): Force is carried on the
inside of the flanges (mode 3), and force must pass (mode 5) through the line between
the flange tips (mode 4), and transverse reinforcement tension (mode 6) and then to the
shear studs on the left-hand side of the column (i.e. sagging side) (mode 1).
d. Load path 4 (LP4, involving Mechanism 2 as shown in Fig. 2d): Force is carried by
mode 3, mode 5, mode 4, and transverse and lateral reinforcement tension (modes 6 and
7) and then to the shear studs on the right-hand side of the column (i.e. hogging side)
(mode 1).
Also, load paths can also act in parallel. If elements of the load path, associated with
the different modes are not strong enough, then strength loss may occur. The strengths
of the different modes are controlled by the actions as shown at the bottom of Fig. 1.
Fig. 2 Load paths for slab internal forces (red: tension, blue: compression)
13
Bulletin of Earthquake Engineering
3 Numerical methodology
The moment frame internal beam-column subassembly with composite deck slab shown
in Fig. 3 was considered based on experimental tests of Chaudhari (Chaudhari 2018).
This subassembly was designed using capacity design principles to obtain strong column/
connection–weak beam behaviour so that the column remains elastic. The tested frame
subassembly was modelled in ABAQUS (version 6.11.2) (Simulia 2011) in 3-D. Beam
(310UB32 Grade 300), column (310UC198 Grade 300), continuity plates, gusset plates,
metal deck, and column continuity plate members, as shown in Fig. 3a, consisted of four-
noded shell finite elements (S4R) (Mago and Clifton 2008). The beam-to-column con-
nection was modelled as rigid, without explicit consideration of the end-plates and bolts
(Chaudhari et al. 2019), this is based on the experimental study conducted at the University
of Canterbury and no damage to end-plate connection, panel zone or column was observed
in any of the tests as these elements were designed to remain elastic. Also, the numeri-
cal model without the joint flexibility (SAP2000 model, (Chaudhari 2018)) gave similar
behaviour to the experimental model.
The concrete slab was modelled using eight noded solid elements (C3D8R),
and the reinforcing steel was modelled using the two noded beam elements (B31).
13
Bulletin of Earthquake Engineering
45 2.5
40
Compressive Stress (MPa)
2.0
20
1.0
15
10
0.5
5
0 0.0
0.000 0.001 0.002 0.003 0.004 0.005 0 0.0005 0.001 0.0015 0.002 0.0025
Compressive Strain
Tensile Strain
Fig. 4 Uniaxial stress–strain curve of concrete under compression and tension loading
Material nonlinearity was incorporated by utilising the nonlinear constitutive law at the
stress–strain level, geometric nonlinearities were assigned using ‘Nlgom’ feature avail-
able in ABAQUS software, and the interface nonlinearity associated with the contact
opening and closing between the column and the slab was simulated using the ‘con-
tact pair’ feature. A displacement control test protocol, with increasing lateral drift, was
applied at the column top.
3.2 Material properties
3.2.1 Concrete
Concrete material was modelled using the concrete damage plasticity (CDP) model
(Simulia 2011). Dilation angle (ψ), flow potential eccentricity (ϵ), ratio of initial biaxial
compressive yield stress to initial uniaxial compressive yield stress (fb0/fc0), and ratio
of second stress invariant on the tensile meridian to the compressive meridian at initial
yield (kc) were considered following Jankowiak and Lodygowski (Jankowiak and Lody-
gowski 2005). In the absence of sufficient relevant tests to identify all parameters, some
values from the literature (Alfarah et al. 2017), reported in Table 1, were used.
Aslani and Jowkarmeimandi (Aslani and Jowkarmeimandi 2012) concrete uniaxial
compression model was used; where the compression envelope is based on Carreira and
Chu (Carreira and Chu 1985), with exponential values for the ascending and descend-
ing branches and compressive stress provided as a tabular function of plastic strain. The
uniaxial compressive stress–strain curve was assumed linear up to 0.4f’c. Thereafter it
was calculated according to Eqs. (1)–(3), as shown in Fig. 4a.
13
Bulletin of Earthquake Engineering
( )
� 𝜀
fc n 𝜀�c
( )n (1)
c
fc =
𝜀
n − 1 + 𝜀c�
c
[ ( )]−0.74 �
n = n1 = 1.02 − 1.17 Esec ∕Ec if 𝜀c ≤ 𝜀c (2)
�
n = n2 = n1 + (a + 28b) if 𝜀c ≥ 𝜀c (3)
3.2.2 Steel
The mechanical properties of the structural steel, such as nominal tensile stress (σnom)
and nominal tensile strain (εnom) were obtained from the tension coupon tests (Chaudhari
2018). Nominal stress and strain properties were converted into true stress (σtrue) and strain
(εptrue) using Eq. 5 and 6 to obtain Fig. 5:
13
Bulletin of Earthquake Engineering
( )
𝜎true = 𝜎nom 1 + 𝜀nom (5)
p ( ) (𝜎 )
𝜀true = ln 1 + 𝜀nom − true (6)
E
The metal deck sheet and the rebars were modelled using a bilinear stress–strain rela-
tionship as shown in Fig. 6 (Liang et al. 2005; Prakash et al. 2011).
For the rebar material, the yield stress was based on the minimum values specified in
AS/NZS:4671 (2001), whereas the yield stress of the metal deck sheet was obtained from
the ComFlor80 (2014) catalogue. The ultimate strength and strain hardening modulus for
the metal deck and rebar was considered equal to 1.28 times yield stress (i.e. σu = 1.28σy)
and 0.0125 times the modulus of elasticity (i.e. Esh = E/80) respectively (Mirza and Uy
2011; Smitha and Kumar 2013). Key properties are reported in Table 2.
The global coordinate system is shown in Fig. 7 with the X-axis representing the frame
subassembly out-of-plane direction. The Y and Z-axes coincide with the longitudinal axes
of column and beam, respectively. The column base was restrained against translation and
rotation, except rotation was permitted about the X-axis. Similarly, to replicate the beam
end support condition of the actual test, rotation was only permitted about the X-axis and
translation along the Z-axis to represent the roller supports. Column top out-of-plane
lateral movement was restrained. These boundary conditions were applied to the master
13
Bulletin of Earthquake Engineering
nodes at the column and beam end centre points. At these sections, to ensure plane sections
remain plane, slave nodes with tie constraints followed the master node to achieve this as
shown in Fig. 7.
3.2.4 Contact elements
3.2.4.1 Slab‑column interaction The strength and stiffness of a composite frame subas-
sembly depends on the degree of composite action between the slab and the beams, as well
as on the robustness of the force transfer loadpaths (e.g. Fig. 2). To simulate the column-
slab interaction, contact surfaces were assigned to the column flanges, web, endplate, gus-
set plates, and relevant slab surfaces. A surface-to-surface contact feature available in the
ABAQUS was used to define the contact interaction between the column and the concrete
slab. The desired interaction was achieved by providing a ‘hard’ contact in the normal direc-
tion (to avoid the penetration into each other) and ‘friction with penalty behaviour’ in the
tangential direction (with a coefficient of friction of 0.2) (Salvatore et al. 2005; Gil and Bayo
2008). Note that the separation between the hard contacts was allowed under uplift. Column
and gusset contact surfaces were modelled as master surfaces, and concrete slab surfaces
were assigned as slave surfaces.
3.2.4.2 Slab‑beam interaction In composite deck slabs, shear studs are provided to trans-
fer longitudinal shear between the steel beam and deck slab thereby limiting the deck slip.
Dowel action is the primary force transfer mechanism where the shear studs must sustain
the slip induced at the steel and concrete element interface (Oehlers and Bradford 1995). In
finite element modelling, the composite behaviour between the steel beam and composite
slab is commonly simulated by either (i) modelling the shear stud as a 3D solid, or beam,
element embedded into the concrete slab, or by (ii) using special elements like connectors/
springs. However, the modelling of the shear studs as a 3D-solid element or beam element
leads to increased computation time as well as difficulties in convergence (Henriques et al.
2013). In contrast, the employment of the spring elements (to simulate the composite action
through the shear studs) is simple to use, gives good convergence, and is computationally
more effective (Smitha and Kumar 2013; Gil and Bayo 2008; Henriques et al. 2013) (Wang
and Tizani 2010). Hence, the non-linear shear springs were provided at each stud location to
represent the stud behaviour in the horizontal and vertical directions.
13
Bulletin of Earthquake Engineering
The beam and the metal deck were modelled with a small gap between the beam top
flange and the deck sheet in order to assign the shear spring. The spring connected the
beam top flange node with the deck sheet node immediately below it at each shear stud
location as shown in Fig. 8c. Beam flanges and deck sheets were modelled using shell
elements with the middle surface as a reference plane. The gap between the beam top
flange and the concrete slab is assumed to be equivalent to the summation of one-half
of the thickness of the beam top flange and one-half of the thickness of the metal deck
sheet (Baskar et al. 2002), as depicted in Fig. 8b. The node-to-node connection between
metal sheet and steel beam top flange is achieved through the non-linear spring element
representing the shear stud as shown in Fig. 8c.
The shear stud force–displacement relationship followed the Johnson and Molen-
stra (Johnson and Molenstra 1991) formulation in Eq. 7 where ‘Prk’ is the characteris-
tic strength of the shear studs of 39kN and 48kN for transverse and longitudinal deck
respectively (Chaudhari 2018; Hicks 2011), and the values for constants ‘α’ & ‘β’ were
selected as 0.989 and 1.535 mm−1 respectively Johnson and Molenstra (Johnson and
Molenstra 1991).
The shear stud non-linear load-slip behaviour shown in Fig. 8d was assigned to the
nonlinear shear spring. The metal deck and beam flange nodes to which the spring con-
nects are constrained against relative movement in the vertical and out-of-plane hor-
izontal directions, and the rotational directions. In terms of decking-beam top flange
(a) Shear Studs in Tested Composite Beam (b) Shear Studs in Finite Element Model
(c) Stud Idealisation as Nonlinear Shear Spring (d) Shear Stud Load-Displacement Behaviour
13
Bulletin of Earthquake Engineering
lateral behaviour, both the slip relation at the studs, and the friction at the other nodes,
contribute.
3.2.4.3 Deck – slab and Rebar‑slab interaction The concrete slab and the metal deck were
assumed to have a perfect bond, without uplift or slip for simplicity. This was realised by
embedding the deck sheet into the concrete slab using the ABAQUS constraint option.
Similar interaction was made for the rebars which were embedded into the concrete slab
(Alashker et al. 2010).
The eight noded solid elements (C3D8R) with the reduced integration points were
adopted to model the concrete slab to capture the concrete local failure (Salvatore et al.
2005; Prakash et al. 2011; Mirza and Uy 2011; Gil and Bayo 2008). The frame sub-
assembly structural elements such as beams, column, gusset plates, column continu-
ity plates and metal deck were modelled using four noded (S4R) shell elements with
reduced integration points (Smitha and Kumar 2013; Alashker et al. 2010; Sadek et al.
2008) using auto-meshing. This resulted in less processing time compared to solid
elements (Kim 2003; Broekaart 2016). Different mesh sizes were modelled to inves-
tigate the sensitivity effect on the overall global response of the frame sub-assembly
with increasing lateral drift. The mesh sensitivity of the shell elements (S4R) was per-
formed on the Bare Steel Frame (BSF) sub-assembly model with three different mesh
sizes 20 mm (BSF-Mono-Mesh 20), 30 mm (BSF-Mono-Mesh 30) and 40 mm (BSF-
Mono-Mesh 40). Whereas the mesh sensitivity of the solid elements (C3D8R) was per-
formed on the Fully Isolated Slab Unit (FISU) sub-assembly model with three different
mesh sizes 50 mm (FISU-Mono-Mesh 50), 75 mm (FISU-Mono-Mesh 75) and 100 mm
(FISU-Mono-Mesh 100). Mesh sizes of 30 mm and 75 mm were selected to eliminate
mesh sensitivity effects for steel (Shell elements, S4R) and concrete (Solid elements,
C3D8R) components respectively, as shown for the frame sub-assembly global force-
drift responses in Fig. 9.
250 250
200 200
150
Force (kN)
150
Force (kN)
100 100
FEM-FISU-Mono-Mesh50
FEM BSF-Mono-Mesh20
FEM-FISU-Mono-Mesh75
50 FEM-BSF Mono-Mesh30 50
FEM-FISU-Mono-Mesh100
FEM-BSF Mono-Mesh40
0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Displacement (mm) Displacement (mm)
(a) Mesh Sensitivity: Shell Element (S4R) (b) Mesh Sensitivity: Solid Element (C3D8R)
13
Bulletin of Earthquake Engineering
Beam ends were provided with vertical roller supports, and the column bottom with a pin
support, to replicate the tested frame subassembly boundary conditions. Beam and column
ends were restrained against out-of-plane displacement. The displacement control test pro-
tocol of Fig. 10 (specified in ACI T1.1-01 (2001)) was applied to the column top as per the
experimental tests.
4 Numerical validation
As part of the experimental research, beam-column subassembly tests with different slab
configurations were conducted at the University of Canterbury (Chaudhari 2018; Chaud-
hari et al. 2019). The four types of tested composite slab configurations are: (i) Fully Iso-
lated Slab Unit (FI-SU), (ii) Shear Key Slab Unit (SK-SU), (iii) Modified Shear Key Unit
(MSK-SU), and (iv) Full Depth Slab Unit (FD-SU). To investigate the contribution of the
composite slab to the over-strength capacity of the beam, a bare steel frame (BSF) sub-
assembly was also tested, which act as benchmark test. Specimen details are summarised in
Table 3 and shown in Fig. 11.
The numerical monotonic force–displacement response and stiffness/strength values
are compared with the experimental results shown in Fig. 12 and Table 4. The numerical
model initial lateral stiffness is up to 17% greater than the experimental values where slip
at the boundary element pins may occur. The strength was generally less than 3% different
but it was almost 9% different for SK-SU. The numerical envelope follows the experimental
hysteresis loop envelope for the BSF, FI-SU, and FD-SU frame sub-assemblies as depicted
in Figs. 12a, b, and e, respectively. However, for the SK-SU frame sub-assemblies, where
the decking was parallel to the main beam, the numerical model overpredicted the strength
because the slab mid-height splitting/delamination failure mechanisms within the column
flange region, which was noted during experimental test (Chaudhari et al. 2019), were not
modelled in FEA.
Figures 13, 14, 15, 16, 17 show that the location and magnitude of the beam flange
buckles compare well with the experiments. The simulated model could capture the buck-
ling of the top and bottom flanges, and the location of the flange buckling in the numerical
model matches with the experimental observations, although the buckle direction, which
depends on the initial imperfections sometimes differed.
13
Bulletin of Earthquake Engineering
13
Bulletin of Earthquake Engineering
Fig. 11 Details of test sub-assemblies. A Bare steel frame (BSF) Test sub-assembly. B Fully isolated slab ▸
unit (FI-SU) Test sub-assembly. C Shear key slab unit (SK-SU) test sub-assembly D Modified shear key
slab unit (MSK-SU) test sub-assembly. E Full depth slab unit (FD-SU) Test sub-assembly
The equivalent plastic strain distribution at 5.0% sub-assembly lateral drift is shown in
Fig. 18, and it can be concluded that the column and the panel zone was in the elastic state,
which was in line with the behaviour observed during the experimental tests conducted at
the University of Canterbury (Chaudhari et al. 2019).
Figures 19, 20, 21, 22 show the minimum principal stress contour plots at the drift
associated with the peak lateral strength from the numerical analysis. The minimum value
represents slab compression. The isolated frame sub-assembly in Fig. 19 indicates low
(< 4 MPa) slab compression because of slab isolation from the column, which deactivated
the force transfer from the column to the slab. This matched the experimental behaviour
(Chaudhari et al. 2019).
The shear key and modified shear key frame sub-assemblies, SK-SU and MSK-SU, were
isolated at the column outer flanges to activate only the force transfer Mechanism 2 without
force transfer Mechanism 1 as described in Fig. 1. As a result, little stress was observed
on the outer side of the flanges as shown in Figs. 20 and 21 and the transfer Mechanism 2
forces are shown. The compression strut angle was approximately 45°, which is consistent
with EN1998-1 (2004). The noted maximum compressive stress in the SK-SU and MSK-
SU frame sub-assemblies was 22 MPa and 34 MPa, respectively, which lies between 0.6fc´
to 0.85fc´ (where fc´ = 40 MPa). This peak stress was locally concentrated at the tip of the
column flanges, indicating the stress concentration on the sagging side of the beam. The
average stress range in compression strut was around 15 MPa (i.e. 0.4fc´). The difference in
the maximum compressive stress between the SK-SU and MSK-SU frame sub-assemblies
was due to the different detailing (i.e. shear key rebars and the confined anchorage).
The compression stress field of the full depth frame sub-assembly in Fig. 22 was differ-
ent to that of the other frame subassemblies because both force transfer mechanisms (i.e.
Mechanism 1 and Mechanism 2) are active. The numerical model shows high stress levels
in front of the column outer flanges (associated with the compression strut of Mechanism
1) and lower stress levels between the column flanges (associated with compression strut of
Mechanism 2). Similar to the observations were made by Salvatore et al. (2005) and Mago
and Clifton (2008). The maximum compression stresses in the Mechanism 1 and Mecha-
nism 2 regions were 28 MPa and 18 MPa, respectively. These indicate 0.7fc´ and 0.45fc,
where fc´ = 40 MPa, which is below the crushing stress, fc´, indicating no strength degrada-
tion from slab spalling is likely.
5 Analytical methodology
For conventional beam-column subassemblies with the slab cast up to the column which
are subjected to lateral force, the beam-column-slab interaction develops internal forces as
shown in Fig. 23. The internal slab force (Ns,L) centroid on the beam sagging side acts at
the topping concrete mid-depth (i.e. tc/2). Whereas on the hogging side, the concrete cracks
and the internal slab force (Ns,R) acts at the rebar mesh elevation. This is the cover to the
longitudinal mesh bars plus one-half of the mesh bar diameter, ɸ/2, from the top of the slab
as shown in Fig. 22a.
13
Bulletin of Earthquake Engineering
(A)
13
Bulletin of Earthquake Engineering
Fig. 12 Comparison of force–displacement of behaviour of numerically simulated result with the test
results
13
Bulletin of Earthquake Engineering
Table 4 Comparison of Sub-assemblies initial stiffness and peak strength of numerical (FEA) model with experimental results
Test specimen Initial stiffness (kN/m) Peak strength (kN)
Numerical (FEA) Experimental Deviation Numerical (FEA) Experimental Deviation
(N-E)/E (%) (N-E)/E
(%)
Bare steel frame (BSF) 6070 6156 − 1.4 203.1 206.0 − 1.4
Fully isolated slab unit (FI-SU) 7590 8747 − 15.2 217.3 211.4 2.7
Shear Key Slab Unit (SK-SU) 12,888 15,031 − 16.6 286.5 263.0 8.1
Modified shear key slab unit (MSK-SU) 13,137 14,506 − 10 288.6 285.2 1.2
Full depth slab unit (FD-SU) 14,136 16,402 − 16 313.7 306.3 2.4
13
Bulletin of Earthquake Engineering
13
Bulletin of Earthquake Engineering
Treating the slab left- and right-hand sides separately, the total slab compression forces
in the steel and concrete on the sagging (left) side, and the tension forces on the hogging
(right) of the frame sub-assembly may be evaluated using the following equations:
where Prk is characteristic strength of the shear stud, nsl and nsr are the number of beam
shear studs in the subassembly on the left and right side of the column respectively, f’c
concrete compression strength, beff is the composite slab effective width, tc is the compos-
ite slab topping thickness), Ag is the steel beam area framing into the column, fy is yield
strength of the steel beam flange, nɸ is number of rebars within the concrete slab effective
13
Bulletin of Earthquake Engineering
width, Aɸ is the individual rebar area of and fyɸ is yield strength of the rebar. The splitting
resistance of the concrete in front of the shear studs was checked as per HERA (Raed et al.
2003) and was found not to govern.
The force transferred from the slab directly to the column outer flanges, Frd1, column
(Mechanism 1) and from the slab to the column sides using the inclined concrete struts,
Frd2, from Mechanism 2 is shown in Fig. 22b, The resistance offered by Mechanism 1 can
be related to the crushing strength of the concrete (Braconi et al. (Braconi et al. 2010))
using Eq. (10) (reproduced from EN1998-1 (EN1998-1 2004)):
�
Frd1 = 0.85fc × Bc × tc (10)
13
Bulletin of Earthquake Engineering
The strength associated with the force transfer Mechanism 2 was developed through the
formation of two compressive struts on the column sides, which is shown in Fig. 22b con-
sidering the inclination angle of 45° (EN1998-1 2004) as shown in Fig. 24.
13
Bulletin of Earthquake Engineering
13
Bulletin of Earthquake Engineering
Mechanism 2 compressive struts are analogous to the inclined struts formed in rein-
forced concrete deep beams in strut-and-tie theory. Column side interaction forces depends
on the (i) bearing force developed on the internal column flanges, (ii) slab shear capacity
between column flange tips, and (iii) compression strut capacity. The compression strut
width (Wstrut) was calculated as:
Wstrut = Hc cos 𝜃 (11)
The compression strut force was calculated using Eq. 12.
�
Fstrut = 𝜐 × 0.85fc × Wstrut × tc (12)
The compression strut’s horizontal and vertical force components (FH and FV) on each
side of the column are given by Eqs. (13) and (14), respectively where υ is the reduction
factor (Plumier and Doneux 2001; EN1998-1 2004), and Wstrut is the compressive strut
width.
( � )
FH = Fstrut cos 𝜃 = 𝜐 × 0.85fc × Wstrut × tc cos 𝜃 (13)
( � )
FV = Fstrut sin 𝜃 = 𝜐 × 0.85fc × Wstrut × tc sin 𝜃 (14)
The shear resistance (Fshear) at the critical section (between column flange tips as shown
in Fig. 23) was calculated using the shear friction method where μ is the friction coefficient
of 1.4 for normal density concrete and N* is the force acting perpendicular to the shear
plane (= FV). Mesh bar shear resistance was not considered here since it may be cut in this
zone for construction ease. However, the shear resistance/capacity at this critical section,
Fshear, may be enhanced by providing shear key rebars (Chaudhari et al. 2019) with area,
Av, yield strength, fyɸ, and α is the angle between shear-friction reinforcement and shear
plane as shown in Eq. (15) (NZS3101:1 2006).
Fshear = Av fy� (𝜇 sin 𝛼 + cos 𝛼) + N ∗ 𝜇 (15)
The internal column flange force resistance on one side of the column Fbearing, is given
in Eq. 16 where Bc is the column flange width, and twc is the column web thickness.
( )
Bc − twc
Fbearing = 0.85fc� × × tc (16)
2
The Mechanism 2 force resistance (Frd2) is governed by the minimum strength of the;
(i) compressive strut horizontal force component, FH , (ii) shear resistance at the critical
section, Fshear, and, (iii) bearing resistance at the internal column flange, Fbearing, as shown
in Eq. 17, where the factor of 2 indicates that the strength may come from both sides of the
web.
⎧2⋅F
⎪ H
Frd2 = min⎨ 2 ⋅ Fshear (17)
⎪ 2 ⋅ Fbearing
⎩
Force transfer Mechanism 3 was developed through the interaction of the shear studs
(installed on the secondary beam) and the concrete. The compression force developed from
13
Bulletin of Earthquake Engineering
the shear studs was transferred to the column through shear and twisting of the transverse
beam (Plumier and Doneux 2001). Figure 25 pictorially explains the slab force transfer
Mechanism 3 from the transverse beam to the column.
The internal force associated with mechanism-3 was calculated based on the recommen-
dations specified by EC8 (EN1998-1 2004), given by Eq. (18).
Frd3 = ns_tr × Prk (18)
where ‘ns_tr’ is a number of shear studs installed within the effective width on the trans-
verse beam. The additional in-plane moment demand on the column panel zone due to
Mechanism 3 is calculated as:
( )
Dbs t
Mrd3 = Frd3 + Ds − c (19)
2 2
At the slab-column interaction, the total interaction force (Fint) developed due to Mecha-
nism 1 (Frd1), Mechanism 2 (Frd2), and Mechanism 3 (Frd3) was calculated as the summa-
tion of the individual forces as given below:
Fint = Frd1 + Frd2 + Frd3 (20)
The governing slab force is a minimum of the internal force developed in the composite
section and the force developed due to the Mechanisms 1, 2 and 3, which is given by the
following equation:
{
Fint
Nslab = minimum of
NsL + NsR (21)
The total moment acting at the column centre (Mcol, CL) accounting for the beam and
slab effect was calculated as:
13
Bulletin of Earthquake Engineering
( ) ( )
Db tc Dc
Mcol,CL = Mpb,L + Mpb,R + Nslab + Ds − + Vb,R + Lp
2 2 2
( ) ( ) (22)
Dc Dbs t
+ Vb,L + Lp + Frd3 + Ds − c
2 2 2
In the above equations, the beam moments (Mpb,L & Mpb,R) are modified to consider the
axial-moment interaction as per NZS3404:Part1:1997 (NZS3404:Part1: 1997), the modi-
fied beam moments for the sagging and hogging sides are calculated assuming that the slab
axial force is applied equally to the steel beams on both sides (with Nslab/2 on each side),
as:
{ ( ( ) )
1.18 1 − Nslab ∕2 ∕Ab,L fy,L Mb,L
Mpb, L = min
Mb,L (23)
{ ( ( / )/ )
1.18 1 − Nslab 2 Ab,R fy,R Mb,R
Mpb, R = min (24)
Mb,R
where Mpb,L and Mpb,R = Modified moment in left (sagging) and right (hogging) beam,
respectively. Nb,L and Nb,R = Axial force in left and right beam respectively, Nb,L = Ns,L and
Nb,R = Ns,R. Ab,L and Ab,R = Cross-sectional area of left and right beam, respectively. fy,L and
fy,R = Steel beam flange yield strength of the left and right beam, respectively. Mb,L and
Mb,R = Beam maximum section moment capacity at the plastic hinge location in left and
right beam respectively (Kawashima 1992; FEMA350, 2000; Bruneau 2011), (fy + fu)/2.
Ze.
The beam shear causing the maximum beam section moment capacity is given by:
Mb,L Mb,R
Vb,L = and Vb,R = (25)
Lb,L Lb,R
where ‘Lb,L’ and ‘Lb,R’ are beam length between the tip of the gusset plate (where the plas-
tic hinge forms), and the point of contraflexure for the left and right beam, respectively.
The predicated maximum lateral strength at the column top (Vcol,prd) is evaluated the story
height (H) as:
Mcol,CL
Vcol,prd = . (26)
H
The lateral strength of the frame subassemblies was estimated using the equations above.
The lateral strength of the frame subassembly corresponding to the yield strength (fy),
average strength ((fy + fu)/2), and the ultimate strength (fu) of the beam was calculated.
Table 5 summarises the subassembly predicted lateral strength and compares it with the
experimental results. The lateral strength corresponding to the yield and ultimate mate-
rial strength of the beam provides the lower and upper bound values. It can be seen from
Table 5 that the lateral strength calculated based on the beam average material strength was
13
13
Table 5 Comparison of predicted lateral strength of frame sub-assemblies with test results
Specimen description Predicted lateral strength (kN) Experimental
lateral strength
Lower bound based on yield Upper bound based on ulti- Average based on strength (kN)
strength (fy) mate strength (fu) (fy + fu)/2
*Test Units TD-SU and LD-SU were both tested by Hobbs (Hobbs 2014). Test LD-SU included a transverse beam with shear studs into the decking
Bulletin of Earthquake Engineering
Bulletin of Earthquake Engineering
in close agreement with the experimental test results. Based on the results summarised in
Table 5, it was concluded that the proposed analytical methodology can reliably predict the
lateral strength of the frame sub-assembly with different slab configurations. It also consid-
ers the strength hierarchy of the failure modes depicted in Fig. 1 and accounts for different
force transfer mechanisms (i.e. Mechanism 1, 2 and 3).
The frame subassemblies considered represent the interior joint of a typical steel frame
building. When the internal frame sub-assembly is subjected to lateral loads, it develops
positive bending (i.e. sagging) in one beam and negative bending (i.e. hogging) in the other
beam. This leads to different second moments of area on the sagging and hogging sides.
Therefore, the subassembly initial lateral stiffness with a composite slab was calculated
using an equivalent moment of inertia (Ieq), which takes into account the effective moment
of inertia on the sagging and hogging sides, which is given by Eq. (27) (EN1998-1 2004;
Cowie 2015):
Ieq = 0.6Ieff _sag + 0.4Ieff _hogg (27)
The effective moment of inertia of the composite beam (Ieff) considering the effect of the
degree composite action was calculated using Eq. (28) (NZS3404:Part1 1997):
( )
Ieff = Ist + 0.85P0.25
comp Itr − Ist (28)
where Ieff_sag and Ieff_hogg = Effective second moment of area of composite beam in posi-
tive (sagging) bending and negative (hogging) bending respectively. Ist = Second moment
of area of steel beam alone. Itr = Composite beam second moment of area transformed into
an equivalent steel section. Pcomp = Degree of composite action.
Comparisons of different subassembly predicted strengths and stiffnesses with the
test results are shown in Fig. 26. Again, it may be seen that the experimental peak lateral
strength falls within the predicted lower and upper bound values. Also, the proposed meth-
odology reasonably predicts subassembly stiffness, as well as strength, for the different slab
configurations but the stiffness does tend to be an upper bound as the model does not con-
sider connection slip.
The column moment demand contribution due to beam and slab deformations is shown
in Fig. 27. It was computed as the flexural strength without the slab effect, divided by the
strength from all terms of the Eq. 26. The presence of the slab contributed around 30–44%
of the column total moment demand, increasing the bare steel frame strength by 62–76%.
For the isolated frame sub-assembly (FI-SU), the moment demands due to the slab were
negligible due to the absence of the active force transfer mechanisms. For the longitudinal
deck frame sub-assembly (LD-SU), all three force transfer mechanisms (1, 2, and 3) were
active, and the moment demands due to the slab was the highest (i.e. around 44%), possibly
due to the effect of Mechanism 3. Note that the steel beam contribution includes the axial
(P)-moment (M) interaction, so the strength associated with the beam on specimens with a
slab is not the same as that without the slab.
While the subassembly lateral strength with the composite deck slab was dictated by a
number of active force transfer mechanisms, the two major parameters affecting subassem-
bly lateral strength were; (i) the degree of composite action and (ii) the effectiveness factor
(ʋ). A parametric study was conducted to evaluate the influence of these parameters on the
subassembly lateral strength with different slab configurations.
The effect of the degree of the composite action (i.e. the number of shear studs) on the
subassembly lateral strength is reported in Table 6. In the full depth frame sub-assembly
(FD-SU), the lateral strength was increased by 10% for the full composite action, from
13
Bulletin of Earthquake Engineering
0 0
-100 -100
Exp-BSF Exp- FISU
-200 Predicted-UB -200 Predicted-UB
Predicted -LB Predicted -LB
-300 -300
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Lateral Drift (%) Lateral Drift (%)
(a) Bare Steel Frame (b) Fully Isolated Slab Unit
300 300
200 200
100 100
0 0
-100 -100
300 300
Lateral Force (kN)
200 200
100 100
0 0
-100 -100
(e) Full Depth Slab Unit (f) Transverse Deck Slab Unit [13]
200
100
0
-100
-200
EXP-LD
-300
Predicted-UB
-400 Predicted -LB
-500
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Lateral Drift (%)
(g) Longitudinal Deck Slab Unit [13]
13
Bulletin of Earthquake Engineering
32.2%
33.0%
34.6%
36.2%
44.1%
80%
70%
60%
100%
50%
40%
67.8%
67.0%
65.4%
63.8%
55.9%
30%
20%
10%
0%
FI-SU SK-SU MSK-FU FD-SU TD-SU LD-SU
the case tested with 46% composite action. However, the lateral strength was limited by
the stud shear strength and the governing force transfer mode (i.e. shear stud strength)
remained the same irrespective of the degree of composite action. This was because the
provision of full depth slab around the column enhances the strengths of the other modes.
On the other hand, for the transverse deck and longitudinal deck frame subassemblies
(TD-SU and LD-SU), the force transfer mode changes from shear stud strength (at 46%
composite action) to the concrete crushing on the column flange (i.e. Modes 2 and 3 in
Fig. 1) with the increase in the degree of composite action. The influence of the degree of
composite action on the lateral strength was seen only in the FD-SU, TD-SU, and LD-SU
frame sub-assemblies, wherein both the force transfer mechanism-1 and 2 were active. In
the case of FI-SU frame sub-assembly were no active force transfer mechanisms, frame
sub-assembly’s lateral strength remains constant irrespective of the increase in the degree
of composite action. Similar observations were made for the SK-SU and MSK-SU frame
sub-assemblies, where only force transfer mechanism-2 was active. It can be concluded
that the lateral strength of the frame sub-assembly depends on the number of active force
transfer mechanisms and the strength associated with the individual component.
The compressive strength of the struts in Mechanism 2 depends on the effectiveness fac-
tor (ʋ) as shown in Fig. 24. It varies from 0.6 to 1.0 (Plumier and Doneux 2001; NZS3101:1
2006; Yun and Ramirez 1996). It’s influence on subassemblies SK-SU and MSK-SU lateral
strength, where only Mechanism 2 was active was about 5.0% when the effectiveness factor
(ʋ) changed from 0.6 to 1.0 as shown in Fig. 28. For the subassemblies with both Mecha-
nisms 1 and 2 active, the effect ʋ on the subassembly strength is negligible. It can be con-
cluded that the effect of the effectiveness factor (ʋ) can be ignored for practical purposes.
6 Conclusions
(i) The 3D finite element model considered the: steel characteristics; concrete char-
acteristics; shear-stud characteristics; slab concrete crushing on column outer, and
13
13
Table 6 Composite action effect (Pcomp) on subassembly lateral strength
Specimen description Degree of composite action Predicted lateral Governing force transfer mode (refer to Fig. 1)
(Pcomp) (%) strength (kN)
250
200
150
100
50
0
FI-SU SK-SU MSK-FU FD-SU TD-SU LD-SU
internal, flanges; the shear resistance between column flange tips; the compression
resistance within the concrete decking; and transverse and longitudinal rebar yield-
ing. The model, calibrated based on the steel and concrete properties and subjected
to monotonic displacements, provided good global estimates of subassembly lateral
force–displacement envelope curves for subassemblies with no slab, a gap between
the concrete and steel column, and with a full depth slab. The numerical model initial
lateral stiffness is up to 17% greater than the experimental values and the strength
was less than 3% to 9%. Frame subassemblies with the concrete slab isolated from
the column external flange, degraded in strength faster in the experiment than in the
model as a result due to the model not capturing all failure modes. The equivalent
plastic strain distribution at 5.0% sub-assembly lateral drift shows that the column
and the panel zone was in the elastic state, which was in line with the behaviour
observed in the experimental tests. The average contact pressure resulted due to
slab-column interaction was found to be lower than the compressive strength of the
concrete (fʹc), and the localised contact pressure was higher than the average contact
pressure, this was due to stress concentration.
(ii) The simple analytical design/assessment methodology, developed using the beam
plastic moment capacity (considering axial-moment interaction) and the slab forces
based on the strut-and-tie mechanism, evaluated the frame subassembly lateral
strength and stiffness. It considered the parameters and deformation modes used for
the FEM approach. Experimental test strength was always within the tolerances con-
sidering different assumptions about the material strengths, and the best estimate of
strength had an accuracy of better than 7% for the comparisons made. Furthermore,
the observed modes of failure were represented. The presence of a slab touching
the concrete column increased the lateral strength by 30% to 44%. Based on the
parametric study, the number of shear studs has negligible effect in case of FI-SU,
and SK-SU and MSK-SU (with only one active mechanism) frame sub-assembly’s
strength, whereas in other frame sub-assemblies (i.e. FD-SU, TD-SU, and LD-SU),
the effect of composite action varies, and it cannot be ignored. Hence the lateral
strength of the frame sub-assembly depends on the number of active force transfer
mechanisms and the strength associated with the individual component.
(iii) The mode of deformation affecting the strength, and failure mechanism varied, based
on the relative properties of the different elements. All modes controlled in some cases.
(iv) For design, the analytical approach developed allows estimation of the key aspects
of subassembly performance. It is fundamental in concept and easy to implement,
encouraging engineers to detail appropriately to maintain load paths. The proposed
13
Bulletin of Earthquake Engineering
analytical model considers the effect of the composite deck slab with various con-
struction detailing, deck orientation as compared to existing analytical model/meth-
odology. It also considers the strength hierarchy of various nonlinear deformation
modes of steel beam plastic hinging, concrete crushing both outside the column
flange and within the beam flanges, slab shear fracture between the column flange
tips, slab longitudinal and lateral reinforcement yielding as it carried the subassem-
bly moments and transferred forces between the steel frame and the slab, and shear
stud deformation. It improves existing NZS3404 methods for behaviour estimation.
Slab effects, when the slab touches the column and shear studs are present, always
cause an increase in demands on the column and panel zone which need to be con-
sidered as part of design. If there are no shear studs, or the column is isolated from
the slab, there is no need to consider the slab effect in the column demands. If the
slab is fully confined around the column, preventing slab degradation during the
large inelastic deformations, then the composite slab strength can be considered to
provide increased beam resistance, possibly allowing smaller steel beams to be used.
Acknowledgements The authors would like to acknowledge the MBIE Natural Hazards Research Platform for
its support to conduct the proposed research study as a part of the Composite Solution Research Project. Addi-
tional support was provided by ComFlor New Zealand, John Jones Steel Research Funding, and Heavy Engi-
neering Educational and Research Foundation (HEERF). All opinions expressed remain those of the authors.
Funding Open Access funding enabled and organized by CAUL and its Member Institutions. No funding
was received to assist with the preparation of this manuscript.
Declarations
Conflict of interest The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License,
which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Com-
mons licence, and indicate if changes were made. The images or other third party material in this article
are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
References
ACI (2001) ACI T1.1-01: Acceptance criteria for moment frames based on structural testing", American
Concrete Institute
Alashker Y et al (2010) Progressive collapse resistance of steel-concrete composite floors. J Struct Eng
136(10):1187–1196
Alfarah B et al (2017) New methodology for calculating damage variables evolution in Plastic Damage
Model for RC structures. Eng Struct 132:70–86
AS/NZS:4671 (2001) "Steel reinforcing materials", Standards New Zealand, Wellington
Aslani F, Jowkarmeimandi R (2012) Stress-strain model for concrete under cyclic loading. Mag Concr Res
64(8):673–685
Baskar K et al (2002) Finite-element analysis of steel-concrete composite plate girder. J Struct Eng-ASCE
128(9):1158–1168
Braconi A et al (2010) Seismic behaviour of beam-to-column partial-strength joints for steel–concrete com-
posite frames. J Constr Steel Res 66(12):1431–1444
13
Bulletin of Earthquake Engineering
Broekaart D (2016) Simuleon FEA Blog - 5 reasons why you should use a mid-surface shell mesh for thin-
walled parts available at "http://info.simuleon.com/blog/5-reasons-why-your-fea-simulations-should-
be-setup-with-a-mid-surface-shell-mesh-for-thin-walled-parts"." (7 June 2017)
Bruneau M et al (2010) Preliminary report on steel building damage from the Darfield earthquake of Sep-
tember 4, 2010. Bull N Z Soc Earthq Eng 43(4):351–359
Bruneau M et al (2011) Ductile design of steel structures. McGraw Hill Professional
Carreira DJ, Chu KH (1985) Stress-strain relationship for plain concrete in compression. J Am Concrete Inst
82(6):797–804
Chaudhari T (2018) Seismic performance evaluation of steel frame building with different composite slab
configurations. Doctor of Philosophy in Civil Engineering Ph.D. Thesis, University of Canterbury,
Christchurch, New Zealand
Chaudhari T et al (2019) Experimental behaviour of steel beam-column subassemblies with different slab
configurations. J Constr Steel Res 162:105699
Chaudhari T et al (2015) Composite slab effects on beam-column subassembly seismic performance.
STESSA 2015, China Architecture and Building Press
Civjan SA et al (2001) Slab effects in SMRF retrofit connection tests. J Struct Eng 127(3):230–237
ComFlor80 (2014) "Product guide ComFlor80, available at "http://www.comflor.co.nz/wp-content/uploads/
ComFlor/Brochures/ComFlor80Brochure.pdf"." (13 January 2014)
Cowie K (2015) Australian/New Zealand Standard for Composite Structures, AS/NZS 2327, Seismic Provi-
sions Development. Steel Innovations Conference 2015, Steel Construction New Zealand
EN1998-1 (2004) Eurocode 8: Design of structures for earthquake resistance - Part 1: General rules, seismic
actions and rules for buildings", European Committee for Standardization, B-1050 Brussels
FEMA350 (2000) Recommended design criteria for new steel moment frame buildings. Federal Emergency
Management Agency, Washington
Gil B, Bayo E (2008) An alternative design for internal and external semi-rigid composite joints. Part II :
Finite element modelling and analytical study. Eng Struct 30(1):232–246
Henriques J et al (2013) Numerical modeling of composite beam to reinforced concrete wall joints Part I :
Calibration of joint components. Eng Struct 52:747–761
Hicks S (2011) Design resistances of 19 mm diameter headed stud connectors through-deck welded within
the ribs of Comflor 60 and Comflor 80 profiled steel decking. HERA, Manukau City, New Zealand,
(Received through private communication from Steve Stickland)
Hobbs M (2014) Effects of slab-column interaction in steel moment resisting frames with steel-concrete
composite floor slabs. Master Thesis, University of Canterbury, Christchurch, New Zealand
Hobbs M et al (2013) Slab column interaction - significant or not? Steel Innovations, SCNZ
Jankowiak T, Lodygowski T (2005) Identification of parameters of concrete damage plasticity constitutive
model. Found Civ Environ Eng 6(1):53–69
Johnson RP, Molenstra N (1991) Partial shear connection composite beams for buildings. Proc Inst Civ Eng
2:679–704
Kawashima K et al (1992) The strength and ductility of steel bridge piers based on loading tests. J Res Jpn, p 29
Kim T (2003) Experimental and analytical performance evaluation of welded steel moment connections to
box or deep w-shape columns." Ph.D. Thesis, University of California, Berkeley
Lee S-J, Lu L-W (1989) Cyclic tests of full-scale composite joint subassemblages. J Struct Eng
115(8):1977–1998
Leon RT et al (1998) Seismic response of composite moment-resisting connections. I: Performance. J Struct
Eng 124(8):868–876
Liang QQ et al (2005) Strength analysis of steel–concrete composite beams in combined bending and shear.
J Struct Eng 131(10):1593–1600
MacRae G, Clifton G (2015) NZ research on steel structures in seismic areas. In: Proceedings of the 8th
international conference on behavior of steel structures in seismic areas, STESSA, Shanghai, China,
pp 44–58
MacRae GA et al (2007) Overstrength effects of slabs on demands on steel moment frames. Pacific struc-
tural steel conference 2007
MacRae G et al (2013) Slab effects on beam-column subassemblies—beam strength and elongation issues.
Compos Construct Steel Concrete VII, pp 77–92
Mago N, Clifton CG (2008) Investigation of the slab participation in moment resisting steel frames (HERA
report R 4–140). New Zealand Heavy Engineering Research Association, Manukau City, New Zealand
Mirza O, Uy B (2011) Behaviour of composite beam-column flush end-plate connection subjected to low-
probability, high-consequence loading. Eng Struct 33:647–662
NZS3101:1 (2006) Concrete structures standard: Part 1 - The design of concrete structures", Standards New
Zealand, Wellington
13
Bulletin of Earthquake Engineering
NZS3404:Part1:1997 (2007) Steel structures standard part 1-Incorporating amendment no 1 and no 2",
Standards New Zealand, Wellington
Oehlers DJ, Bradford MA (1995) Composite steel and concrete structural members : fundamental behav-
iour. Elsevier, Kidlington, Oxford
Plumier A, Doneux C (2001) Seismic behaviour and design of composite steel concrete structures.
ECOEST2 and ICONS, LNEC Lisboa, Portugal
Prakash A et al (2011) Three dimensional FE model of stud connected steel-concrete composite girders sub-
jected to monotonic loading. Int J Mech Appl 1(1):1–11
Raed Z et al (2003) Shear Stud Capacity in Profiled Steel Decks. HERA Steel Design and Construction Bul-
letin (DCB No. 76)
Sadek F et al (2008) Robustness of composite floor systems with shear connections: modeling, simulation,
and evaluation. J Struct Eng 134(11):1717–1725
Salvatore W et al (2005) Design, testing and analysis of high ductile partial-strength steel–concrete compos-
ite beam-to-column joints. Comput Struct 83(28):2334–2352
Simulia (2011) ABAQUS version 6.11. User’s manual." Dassault Systemes
Smitha MS, Kumar SRS (2013) Steel-concrete composite flange plate connections - finite element modeling
and parametric studies. J Constr Steel Res 82:164–176
Umarani C, MacRae G (2007) A new concept for consideration of slab effects on building seismic perfor-
mance. J Struct Eng Struct Eng Res Centre Chennai India 34–3:34
Wang Z, Tizani W (2010) Modelling techniques of composite joints under cyclic loading. In: Computing in
civil and building engineering, proceedings of the international conference, (Nottingham University
Press, Nottingham, UK), Paper 254, p 507
Webb G et al (2018) Column moment demands from orthogonal beam twisting. Proc Key Eng Mater Trans
Tech Publ, pp 259–269
Yamada S et al (2009) Full scale shaking table collapse experiment on 4-story steel moment frame: Part 1
outline of the experiment. Behav Steel Struct Seismic Areas: STESSA 2009:125
Yun YM, Ramirez JA (1996) Strength of struts and nodes in strut-tie model. J Struct Eng 122(1):20–29
Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
13