Sei sulla pagina 1di 12

Strength of Circular Concrete-Filled Tubes

with and without Internal Reinforcement


under Combined Loading
Jiho Moon 1; Dawn E. Lehman, M.ASCE 2; Charles W. Roeder, M.ASCE 3;
and Hak-Eun Lee 4
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Concrete-filled tubes (CFTs) have been used in civil engineering practice as piles, caissons, columns, and bridge piers. Relative to
conventional structural steel and reinforced concrete components, CFTs have several advantages. The steel tube serves as both reinforcement
and formwork, eliminating the need for both, and provides large tensile and compressive capacities; the concrete fill restrains buckling of the
steel tube, which increases the strength, stiffness, and deformability of the section. In some cases, internal reinforcement is used to enhance
the strength and facilitate connection to adjacent members. Although these properties are well accepted, the use of CFTs in practice is
awkward because design provisions among codes vary significantly and previous research has not considered internal reinforcement.
An analytical research study was undertaken to evaluate and improve design provisions for CFTs with and without internal reinforcement
under combined axial load and bending. A continuum model was developed to simulate prior test results subjected to combined loading
and the validated model was used to investigate the strength and inelastic performance of CFTs under combined loading. Current design
provisions for CFTs were evaluated using the results of these finite-element analysis and previous test results. The comparisons indicate
that current design approach provides good prediction of CFT capacity subjected only to bending or axial demands, but current provisions
provide conservative values for the CFTs under general combined loading. An alternative P-M interaction curve for CFTs was proposed.
DOI: 10.1061/(ASCE)ST.1943-541X.0000788. © 2013 American Society of Civil Engineers.
Author keywords: Concrete-filled tubes (CFTs); Composite behavior; P-M interaction curve; Structural design; Buckling; Nonlinear
finite-element analysis; Metal and composite structures.

Introduction concrete fill is not necessary, however, internal reinforcement is


occasionally used in piles, caissons, and bridge piers in order to
Concrete-filled tubes (CFTs) are composite members that consist of enhance the strength and facilitate connection to other elements.
a steel tube and concrete infill. In Asia, they have been used as CFTs have had limited use in the United States. In part, this
building columns and bridge piers as an alternative to conventional results from U.S. construction practices and inconsistencies in
reinforced concrete construction. A primary benefit of CFTs is the design approach. Currently, AISC (2010), American Concrete
that the concrete fill provides large compressive load capacity Institute (ACI) (2008), and AASHTO (2009) provide design pro-
and restrains buckling of the steel tube, which results in signifi-
visions for CFTs. AISC permits two methods for predicting the
cant resistance and inelastic deformation capacity (Marson and
resistance of CFT members: (1) a simplified plastic stress distribu-
Bruneau 2004; Roeder et al. 2010). They also provide economy
tion approach, or (2) a more complex strain compatibility method.
and rapid construction because the steel tube replaces the form-
work and reinforcement. The use of self-consolidating concrete Similarly, ACI uses a general strain compatibility method that
without vibration of concrete further accelerates construction can be simplified to a plastic stress distribution similar to AISC.
(Roeder et al. 2010). In general, internal reinforcement in the The three major differences in the simplified approaches are
that (1) AISC approximates the stress in the uniform block to
1
Research Professor, Civil, Environmental & Architectural Engineer- be as 0.95fc0 for CFTs while ACI approximates the uniform
ing, Korea Univ., 335 Engineering Building, Seoul 136-701, South Korea. strength as 0.85fc0 , (2) AISC uses the entire height of the stress
E-mail: jmoon1979@gmail.com
2
block while ACI recommends that it be reduced by a factor β
Associate Professor, Dept. of Civil & Environmental Engineering, (β < 0.85 depending on concrete strength), and (3) ACI limits the
Univ. of Washington, 214B More Hall, Seattle, WA 98195. E-mail:
delehman@u.washington.edu axial strength to 80% of the squash load with an additional factor to
3
Professor, Dept. of Civil & Environmental Engineering, Univ. of account for eccentricity where AISC treats the instability explicitly.
Washington, 233B More Hall, Seattle, WA 98195 (corresponding author). AASHTO also addresses the design of circular CFTs, but these pro-
E-mail: croeder@u.washington.edu visions are less mature than the AISC and ACI provisions
4
Professor, Civil, Environmental & Architectural Engineering, Korea AISC and ACI also differ on column design in their handling
Univ., 305 Engineering Building, Seoul 136-701, South Korea. E-mail: of both length effects and effective flexural stiffness with respect
helee@korea.ac.kr
to long-term effects for gravity loads. Several researchers (Varma
Note. This manuscript was submitted on August 18, 2011; approved on
December 7, 2012; published online on December 10, 2012. Discussion et al. 2002; Bruneau and Marson 2004; Choi et al. 2006; Roeder
period open until February 9, 2014; separate discussions must be submitted et al. 2010) have investigated the design method of CFTs, yet
for individual papers. This paper is part of the Journal of Structural En- there is not consensus on the correct approach among the different
gineering, © ASCE, ISSN 0733-9445/04013012(12)/$25.00. design provisions.

© ASCE 04013012-1 J. Struct. Eng.

J. Struct. Eng. 2013.139.


In addition, CFT sections may have internal reinforcement and but concrete stresses of 0.95fc0 and 1.0fc0 are used in AISC
there are no test data or rigorous analyses to support the design and Eurocode 4, respectively. The coefficient of 0.95 was used
methods for a reinforced CFT component. This research project in this paper.
was undertaken to develop a uniform and consistent approach to In Fig. 1, t is wall thickness of the tube, y is the distance from the
designing CFT beam-column elements (i.e., elements subjected center of the section to the neutral axis for the specific load combi-
to combined loading) with or without internal reinforcement. Veri- nation, rm is the distance from the origin to the center of the steel
fied nonlinear continuum models were used where the verification tube, ri is the radius from the origin to the inside of the steel tube,
included both CFT components subjected to axial load only or c is the half-width of the compression area, fy and fyb are yield
combined loading. The nonlinear continuum model directly simu- stress of the steel tube and reinforcement, respectively, and fc0 is
lated confinement effects, local buckling, bond stress, and slip be- the compressive strength of the concrete. The variable P is the
tween steel tube and concrete infill, thereby permitting a robust applied axial force and P is positive when it acts in compression.
parametric study. For simplicity, the internal reinforcement was replaced by an inter-
The verified model was used to conduct a parametric study to nal steel ring with an equivalent total area and an equivalent thick-
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

investigate the impact of diameter to thickness (D=t), internal ness, teq , and radius, rbm .
reinforcement (ρi ), and axial load (P=Po ) ratios. The results were Based on the equilibrium condition in Fig. 1, the moment, M,
combined with past experimental results to evaluate internationally and axial force, P, can be determined from an assumed neutral axis
accepted design provisions including AISC (2010) and Eurocode 4 depth, as provided in Eqs. (1a) and (1b)
[European Committee for Sandardization (CEN) 2004] provisions.
Based on the results, an alternative P-M interaction curve for θ
P ¼ −2fyb NAb b − 4fy trm θ
CFT was proposed. The proposed curve is based on the plastic- π
  
stress distribution method (PSDM) combined with stability expres- 0.95fc0 2θ
sions to estimate the flexural strength of low and moderate axial þ ðπ − 2θÞr2i − 2cy − NAb 1 − b ð1aÞ
2 π
load ratios.

2 r2
Design Approach for Circular CFTs with and without M¼ fyb NAb rbm cos θb þ 4fy tc m
π ri
Internal Reinforcing   
0 2c 2 2 rbm cos θb
þ 0.95f c ðr − y Þ − NAb ð1bÞ
To assess the flexural strength of CFTs under combined loading, 3 i π
either the plastic-stress distribution (AISC 2010; CEN 2004) or
train-compatibility methods (AISC 2010; ACI 2008) are used. where
The plastic stress distribution method in AISC (2010) was adapted y y
by Leon et al. (2007) from the Eurocode (Roik and Bergmann sin θb ¼ ; sin θ ¼ ;
rbm rm
1992). Roeder et al. (2010) reported that the plastic-stress distribu- ð1cÞ
tion method is the method of choice for its simplicity and accuracy. NAb
c ¼ ri cos θ; and teq ¼
Thus, only the plastic-stress distribution method was considered in 2πrbm
this study.
The plastic-stress distribution method is an equilibrium-based In Eq. (1), N and Ab are the number of internal reinforcing bars
method used to determine the strength of a CFT member where and the individual bar area, respectively. Eq. (1) is valid when the
instability effects due to global buckling are not considered. Fig. 1 value of θb is between −π=2 and π=2.
shows a cross section of a CFT component with internal reinforce- Fig. 2(a) shows implementation of the PSDM (dashed line) for a
ment. The figure also shows a typical stress distribution in the tube, CFT cross section without internal reinforcement and a D=t ratio of
reinforcement, and concrete fill. The length of each of the stress 60 (therefore, ρi ¼ 0). The axial load and moment are normalized
components is determined by satisfying equilibrium for a given bending only and axial load only cases, where Po;AISC and M o;AISC
externally applied axial load and moment. The AISC (2010) and are the squash load (M ¼ 0) and flexural strength corresponding to
Eurocode 4 (CEN 2004) plastic distribution methods are similar, no axial load (P ¼ 0) from AISC provisions, respectively.

Asc: compression area of steel tube, Acc: compression area of concrete infill
Abc: compression area of reinforcement, Ast: tension area of steel tube
Abt: tension area of reinforcement
2C
Steel tube Concrete infill
fy fyb 0.95fc fyAsc
0.95fc Acc
N.A.
fybAbc
rbm rm y
t b
teq o
P
ri fyAst
fybAbt
fy fyb
Internal Steel Rebar Concrete Equilibrium
reinforcement stress stress stress condition

Fig. 1. Geometry of CFTs with internal reinforcement, stress distribution, and equilibrium condition

© ASCE 04013012-2 J. Struct. Eng.

J. Struct. Eng. 2013.139.


1.2 1.2 D/t=60, L/D=8, i=0% 1.2 D/t=60, L/D=8,
D/t=60, L/D=8, i=0% i=0%
A A
1.0 1.0 1.0 A
Interaction curve Interaction curve
from PSDM A' A"
0.8 A' 0.8 0.8
A' from PSDM
P
P/Po,AISC

P/Po,EC4
k P

P/Po,EC4
0.6 C H 0.6 0.6 H
AISC bilinear curve C'
0.4 0.4 0.4
without resistance factor d
D Interaction curve
0.2 D' 0.2 0.2 including global buckling
E
B B B
0.0 0.0 0.0
0.0 0.5 1.0 1.5 0.0 0.5 1.0 1.5 0.0 0.5 1.0 1.5
M/Mo,AISC M/Mo,EC4 M/Mo,EC4
(a) (b)
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Example of P-M interaction curves: (a) AISC; (b) Eurocode 4

This interaction curve does not account for buckling. To con- factor on the gross cross-sectional stiffness of the concrete to ac-
sider length slenderness effect, the critical buckling load, Pcr , count for material nonlinearity (e.g., cracking).
must be determined. Based on the AISC provisions, Pcr can be Three different values of C3 were investigated in this study [EIeff
determined as from AISC (2010), Eurocode 4 (2004), and Roeder et al. (2010)]
2
and are given by
Pcr =Po;AISC ¼ 0.658λ when λ ≤ 1.5 ð2aÞ  
As
C3 ¼ 0.6 þ 2 ≤ 0.9 by AISC ð2010Þ ð5aÞ
Pcr =Po;AISC ¼ 0.877=λ2 when λ > 1.5 ð2bÞ As þ Ac

where C3 ¼ 0.6 by Eurocode 4 ðCEN 2004Þ ð5bÞ


sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Po;AISC πEIeff  
λ¼ ; Pecr ¼ ; and C3 ¼ 0.15 þ
P
þ
As
≤ 0.9 by Roeder et al: ð2010Þ ð5cÞ
Pecr ðkLÞ2 Po As þ Ac
Po;AISC ¼ 0.95f c0 Ac þ f y As þ f yb Ar ð2cÞ

The variables used in Eq. (2) are defined as follows: k is the After the P-M interaction curve from the plastic-stress distribu-
effective buckling length factor; EIeff is the effective stiffness; tion method and Pcr are determined, the P-M interaction curve
Ac and As are the area of the concrete infill and steel tube, respec- considering the effect of global buckling can be obtained. Both
tively; and Ar is the total area of the internal reinforcement. AISC (2010) and Eurocode 4 (2004) provide methods.
Similarly, Eurocode 4 (CEN 2004) also provides the equation AISC provides three different methods to construct the P-M
for Pcr as interaction curve, and Method 2 was adopted in this study because
this method is based on the plastic-stress distribution method.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fig. 2(a) highlights four points used to establish the P-M interac-
Pcr =Po;EC4 ¼ 1=ðϕ þ ϕ2 − λ2 Þ ≤ 1 ð3aÞ
tion surface based on Method 2 in AISC provisions (AISC 2010):
point A is the full squash load (M ¼ 0), point B corresponds to
ϕ ¼ 0.5½1 þ αðλ − 0.2Þ þ λ2 ; and flexural strength without axial load (M o;AISC ), point C has the same
flexural strength as point B but with axial load, and point D cor-
Po;EC4 ¼ fc0 Ac þ f y As þ fyb Ar ð3bÞ responds to an axial load of one-half of that determined for point C.
The AISC method reduces points A, C, and D to points A 0 , C 0 , and
where α = imperfection factor, which accounts for initial imperfec-
D 0 to account for global buckling effect by multiplying by a reduc-
tion and residual stress of the section. The values of α are 0.21 and
tion factor of the critical load to the squash load Pcr =Po;AISC .
0.34 for buckling curves a and b, respectively, which depend on the
However, with this approach, the reduced flexural strength may
section properties of CFTs. Eurocode 4 (CEN 2004) accounts for
not be conservative at point D 0 because it may result in larger flexu-
the confinement effect of the CFT in predicting the squash load
ral strength than predicted by the plastic-stress distribution method.
when λ < 0.5 and el =D < 0.1, where el is the eccentricity of the
This potential problem can be eliminated by (1) estimation of D 0 on
load. Thus, Po;EC4 could be larger than that from Eq. (3b). The de-
the PSDM curve or (2) removal of point D from the interaction
tailed equations are shown in Eurocode 4 (CEN 2004).
surface; the latter approach is adopted by AISC (2010). The result-
The effective stiffness of the CFT is an important section prop-
ing simple design interaction curve for CFTs is then constructed by
erty that is needed to determine the buckling capacity. Generally,
connecting points A 0 , C 0 , and B, as shown in Fig. 2(a). This bilinear
the effective stiffness of the CFT is expressed in the form
interaction curve is used through this paper for comparison with
EIeff ¼ Es I s þ Eb I b þ C3 Ec I c ð4Þ results of analysis and tests.
Fig. 2(b) shows the construction of the P-M interaction curve
where Es , Eb , and Ec = modulus of elasticity values of the steel including the global buckling effect based on Eurocode 4 (CEN
tube, internal reinforcement, and concrete infill, respectively; I s , 2004). Points A and B [Fig. 2(b)] are the same as points A and
I b , and I c = second moment of area of the steel tube, internal B in AISC (2010). Similarly to AISC, the full squash load is re-
reinforcement, and concrete infill, respectively; and C3 = reduction duced to point A 0 with reduction factor Pcr =Po;EC4. The moment

© ASCE 04013012-3 J. Struct. Eng.

J. Struct. Eng. 2013.139.


capacity at point A 0 , μk , is then removed from the interaction curve relative movement (slip) between the internal reinforcement and
and this removal moment is assumed to decrease linearly to the concrete was not specifically modeled; the truss elements sim-
the point E. The detail of determination of point E is shown in ulating the reinforcement were assumed to be perfectly bonded to
Eurocode 4 (CEN 2004). the concrete and were modeled using the EMBEDED option in
ABAQUS. Similar modeling approaches have been proposed by
previous researchers (Elremaily and Azizinamini 2001; Leon and
Analytical Model Development and Verification Hu 2008).
Boundary and loading conditions used are shown in Fig. 3(b).
The nodes at the base of the CFT column model were fully re-
Model Overview
strained to model a fixed-base column. Distributed axial loads were
A total of 197 nonlinear finite-element models were analyzed with applied uniformly at the top of the model to both the concrete and
the ABAQUS nonlinear finite-element analysis program to investi- steel section. A monotonic lateral displacement history was applied
gate the strength of the CFT (36 models for CFTs under bending, at the top of the model. To ensure convergence, a mesh refinement
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

48 models for CFTs under axial load, and 113 models of CFTs study was conducted. Based on this study, a total of 20 elements is
under combined loading). Fig. 3 shows the configuration of the recommended to model the steel tube. This mesh was adopted for
finite-element model used to model a cantilever columns subjected analysis as shown in Fig. 3(a).
to combined loading. The model must accurately simulate the full The concrete was modeled using the available damaged plastic-
response to axial and bending loads, including any relative move- ity model incorporated in ABAQUS (Lubliner et al. 1989; Lee and
ment between the steel shell and concrete fill. To enable this, four Fenves 1998). This constitutive model simulates triaxiality depen-
different types of elements were used to model the concrete fill, dent plastic hardening. The uniaxial unconfined stress-strain rela-
steel shell, internal reinforcement, and interface between the steel tionship proposed by Saenz (1964) was employed in which it was
shell and concrete fill, as shown in Fig. 3(a). The eight-node solid assumed that the compressive stress-strain relationship is linear up
element (C3D8R) was used to model the concrete, the four-node to a stress of 0.5fc0 and the maximum compressive strength, f c0 , is
shell element (S4R) was used to model the steel shell, and the two- achieved when compressive strain is 0.003. The modulus of elas-
node truss element (T3D2) was used for the internal reinforcement. ticity of the concrete Ec was approximated as 4,700ðf c0 Þ0.5 MPa
The GAP element in ABAQUS was used to simulate the inter- (ACI 2011).
face between concrete infill and steel tube. The GAP element has The tensile response of the concrete was modeled as follows:
infinite stiffness for the compression and no stiffness in tension, (1) the tensile stress increases linearly up to maximum tensile
which permitted simulation of slip by separation of two nodes. strength of the concrete, which is approximate as 0.09fc0 , and
(Penetration of one node into an adjacent one was prevented.) (2) after this peak, the tensile stresses decrease linearly to zero
The normal stresses in the GAP element results in confinement at a strain equal to 10 times the strain at maximum tensile strength
of the concrete, thus it permits explicit modeling of the confining of the concrete, εct .
effect of the tube. Shear stress transfer between the steel tube and The dilation angle, ψ, of the concrete is an important model
the concrete infill is accomplished through friction, which was in- parameter, which in part determines the plastic hardening of the
troduced to the model through a friction coefficient assigned to the concrete. It was approximated as 20° based on the results of prior
GAP element. Thus, shear stress at the interface is generated with research and parametric studies (Moon et al. 2012). For the steel
pressure acting through the GAP element with a specified value tube and internal reinforcement, a trilinear stress-strain relationship
of friction coefficient. A friction coefficient of 0.47 was used based was used with an isotropic-hardening plasticity rule. Young’s
on validation from a prior study (Moon et al. 2012). In contrast, modulus, Es , was approximated as 200,000 MPa and Poisson’s

Concrete infill: Steel tube:


8-node solid element 4-node shell element
Fixed end: U1, 2, & 3 of concrete
infill & steel tube are restrained.

CFT with internal rebar

Step 2: U2 is
applied on the
steel tube. U2
Interface between
Internal rebar:
steel tube & concrete infill:
2-node truss element
GAP element
U3 U1

Step 1: Axial load (Direction 3) is applied


to both concrete infill & steel tube

(a) (b)

Fig. 3. Finite-element model of cantilever CFT specimen

© ASCE 04013012-4 J. Struct. Eng.

J. Struct. Eng. 2013.139.


ratio, υs , was approximated as 0.3. The plastic plateau terminated from 8 to 30 for Matsui et al. (1995). These test data were selected
when strain of the steel, εs , was set equal to 10 times of yield strain for the completeness of the available information.
of the steel (10εsy ) and stress increased up to ultimate strength of The test specimens were simulated to verify the finite-element
the steel, fu , which is achieved when the strain of the steel, εsu , is model used in this study. A half-sine wave initial imperfection was
0.1. The measured stress-strain curve from the CFT tests was used introduced to initiate global buckling. Two different imperfections
to verify the analytical models when available. were examined, including maximum amplitudes, e, of L=2,000 or
L=5,000. Figs. 4(a and b) show the comparison of simulated axial
load-transverse displacement to the test results for Specimen 18-0
Model Verification: CFT under Bending
from Matsui et al. (1995) and Specimen SC154-4 from Han
The bending behavior of CFTs was verified in a prior study (Moon and Yan (2000). The simulated results show good agreement with
et al. 2012). In this paper, the verification was extended to include test results.
slender CFTs under axial load and CFTs under axial-bending com- Analysis results with an eccentricity, e, of L=2,000 provided
bined loading. The brief summary of verification of bending behav- better simulation of tests, as shown in Table 1. The differences be-
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

ior is as follows. tween the simulated and measured test results are likely due to sev-
Three-point bending test results conducted by Thody (2006) eral factors. First, it is almost impossible to achieve ideal pin-ended
were used to verify the finite-element model for the CFTs under support in experiments due to the friction generated between the
bending. The tests were simply supported components with a test specimen and the test setup. Second, the exact shape and mag-
5.49-m span and a single actuator located at midspan to apply nitude of initial imperfection are unknown and cannot be applied to
the cyclic lateral load. The steel tubes used for the test specimens the finite-element models. These factors indicate the finite-element
were 508 and 6.35 mm in diameter and thickness, respectively. The models used in this study provide a suitable, conservative estimate
measured compressive strength of the concrete, f c0 , and yield stress of the buckling strength of CFTs.
of the steel tube, f y , were 84.1 and 520.9 MPa, respectively.
The simulated moment-drift and slip-drift relationships agree
Model Verification: CFT under Combined Loading
well with those from the test. On average, the analysis estimated
a bending capacity approximately 6% smaller than that achieved The combined axial and bending test results of Marson and
in experiments. The computed maximum slip was slightly larger Bruneau (2004) were used to verify the finite-element analysis
than the experimentally measured slip. The local buckling shape models for CFT under general combined loading. The tests were
and cracking pattern that was observed by removing the steel tube conducted using a footing that provided near full fixity; the
after the test were also compared with simulation results, and it was CFT beam columns were connected to a pair of wide flange sec-
found that analysis results agree well with the test results. tions and embedded in the footing. As a result, the footing was not
modeled, but a fixed-base column model was employed. Table 2
provides geometry and material properties for the four specimens
Model Verification: CFT under Axial Load
simulated. The D=t ratio varied from 43.2 to 73.9 and axial load
Han and Yan (2000) and Matsui et al. (1995) conducted experi- ratio (P=Po;AISC ) ranged from 0.13 to 0.32.
ments on the buckling strength of a pin-ended CFT column. A total Fig. 5 compares the measured and simulated moment-drift re-
of 15 test specimens were analyzed. The profiles of their specimens lationship where the drift is defined as lateral displacement at the
and comparison results are shown in Table 1. The D=t ratios were top of a specimen divided by the column length. The models pro-
24 and 39.6 for the specimens of Han and Yan (2000) and Matsui vide good simulation of the measured stiffness, yield, and maxi-
et al. (1995), respectively. The ratios of the length to the diameter, mum strengths. The average difference in the predicted and
L=D, varied between 32.5 and 38.5 for Han and Yan (2000) and measured strengths was 7.3%. In the test program, cyclic loading

Table 1. Specimen Properties and Predicted Results for CFT under Axial Load
Average Average error Number of
Researcher D (mm) D=t L=D f y (MPa) fc0 (MPa) error (%) (e ¼ L=2,000) (%) (e ¼ L=5,000) specimens
Matsui et al. (1995) 165.2 39.6 12–30 358.7 40.9 2.8 9.5 4
Han and Yan (2000) 108 24 32.5–38.5 348.1 25.4–37.4 13 12.5 11

1600 18-0, Matsui et al. (1995), Test 400 SC154-4, Han & Yan (2000), Test
FEM, e=L/2,000
Applied axial load P

FEM, e=L/2,000
Applied axial load P

1200 P 300 P

800 200
L L
400 100

0 0
0 20 40 60 0 50 100 150 200
(a) Displacement at midspan (b) Displacement at midspan

Fig. 4. Verification of the finite-elemeny model for CFTs under axial load: (a) 18-0 specimen (data from Matsui et al. 1995); (b) SC154-4 specimen
(data from Han and Yan 2000)

© ASCE 04013012-5 J. Struct. Eng.

J. Struct. Eng. 2013.139.


Table 2. Specimen Properties and Predicted Results for CFT under The CFT component was modeled as a cantilever to easily evaluate
Combined Loading combined loading. The ratio of the length to the diameter of the
fy f c0 Mu;Test = CFT L=D was 8 for the bending and reference models. This
Model D (mm) D=t L=D (MPa) (MPa) P=Po;AISC M u;FEM L=D ratio is equivalent to 16 for simply supported CFTs under cen-
CFT34 323.9 43.2 6.8 415 40 0.32 1.07
tral concentrated loads. According to a test program conducted by
CFT42 406.4 42.8 5.4 505 35 0.19 0.96 previous researchers (Elchalakani et al. 2001; Han et al. 2006;
CFT51 323.9 58.9 6.8 405 35 0.33 1.15 Thody 2006) for which L=D ratio of CFTs under bending varied
CFT64 406.4 73.9 5.4 449 37 0.13 0.98 from 6 to 15, an L=D of 16 minimized the effect of shear stress on
the bending behavior of CFTs. The diameter of the CFT was
1,524 mm (60 in.), which represents the typical diameter of the
bridge pier. Yield stress of the steel, f y , and compressive strength
was applied and test specimens lost lateral strength after tearing of of the concrete, fc0 , were 344.5 MPa (50 ksi) and 34.5 MPa (5 ksi),
steel tube followed local buckling; the tearing of a steel tube was respectively, for all models. The yield stress of internal reinforce-
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

not modeled and therefore the impact on the postpeak response is ment, fyb , is 344.5 MPa (50 ksi).
not simulated. Thus, postpeak behavior at high drift ratio obtained The parameters were studied as follows. The D=t ratios studied
from the analysis could be different from that of the test. were 40, 60, 80, and 100. The internal reinforcement ratio, ρi , was
Fig. 6 shows the comparison of local buckling shape, stress dis- varied and included 0.0, 0.5, 1.0, 2.0, and 3.0%. Two different con-
tribution, and cracking pattern of the CFT64 specimen. The ob- crete cover depths, dc , of 50.8 mm (2 in.) and 152.4 mm (6 in.)
served location and shape of the local buckling is similar to the were used resulting in radii ratios, rbm =rm , of 0.933 and 0.8, respec-
simulated response, as shown in Fig. 6. The steel tube fully yielded tively (Fig. 1). The resulting values of the buckling parameter λ are
in tension and compression region and the maximum compression summarized in Table. 3.
was observed at the tip of the compression part. Further, at 7% drift,
cracking in both tension and compression side was observed, which
was also well simulated. Bending with No Axial Load
The flexural strength of the CFT is needed to establish P-M inter-
action curves, and 36 CFT members were analyzed as a fixed can-
Parametric Study of CFT under Combined Loading tilever column and the lateral displacement was applied at the top
of the CFT. Fig. 7 summarizes these analytical results. Fig. 7(a)
The verified models were used to study the salient parameters shows the variation of moment capacity normalized by the moment
including amount of internal reinforcement, D=t ratio, axial load capacity of the CFT without internal reinforcement as a function of
ratio, and relative radius of the reinforcement and the steel tube. ρi =ρtotal , where ρtotal is the total reinforcement ratio including steel
Finite-element analyses were performed on CFT models with tube and internal reinforcement.
and without internal reinforcement to establish P-M interaction The moment capacity increased with increasing ρi =ρtotal . The
curves, which include the effects of global and local buckling. flexural strength of the CFT increased by approximately 50% when

600 Test results, CFT34 1200 Test results, CFT42


FEM results, CFT34 FEM results, CFT42
Base Moment (kN. m)

Base Moment (kN. m)

500 1000

400 800
P P
300 H 600 H

200 400
L L
100 200

0 0
0 2 4 6 8 10 0 2 4 6 8 10
(a) Drift (%) (b) Drift (%)

500 Test results, CFT51 800 Test results, CFT64


FEM results, CFT51 FEM results, CFT64
Base Moment (kN. m)

Base Moment (kN. m)

400
600

300
P P
H 400 H
200
L 200 L
100

0 0
0 2 4 6 8 10 0 2 4 6 8 10
(c) Drift (%) (d) Drift (%)

Fig. 5. Verification of the finite-element model for CFTs under combined loading: (a) CFT34; (b) CFT42; (c) CFT 51; (d) CFT64 specimen

© ASCE 04013012-6 J. Struct. Eng.

J. Struct. Eng. 2013.139.


* Von-Mises stress in steel tube

Yielding
* Min. Principle stress (Compression) in concrete infill
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

* Comparison of local buckling


shape (at 7% drift)

* Cracking pattern of concrete infill


Max.
compression

Load direction

Fig. 6. Comparison of local buckling shape, stress distribution, and cracking pattern of CFT64 specimen

Table 3. Range of D=t Ratio and λ Fig. 7(d) shows the comparison of the predicted strength from
λ [based on Number current design codes with previous test results for the CFT without
Loading case D=t ratio AISC (2010)] of models internal reinforcement (Prion and Boehme 1994; Elchalakani et al.
2001; Han et al. 2006; Thody 2006). Again, AISC (2010) and
Flexure
Eurocode 4 (2004) provisions provide the conservative strength
FEM 40–100 — 36
Test data 40.2–109 — 33
estimates of CFTs under bending; the average discrepancy be-
Pure axial load tween test and predicted strength was 23.2 and 22.9% for AISC
FEM 40–100 0.70–1.9 48 and Eurocode 4 provisions, respectively. The scatter is compara-
Test data 18.4–68.7 0.34–2.17 113 tively large for the measured-to-code predicted response, as shown
Combined loading in Fig. 7(d). This is mainly attributed to the uncertainty of material
FEM 40–100 0.70–1.14 113 properties of test specimen or test procedure. In the case of the
Test data 34.5–111.9 0.31–0.65 34 analysis, the f u =f y ratio is the same for all models and scatter is
quite uniform. On the other hand, for the actual test, the fu =f y ratio
is not constant. Furthermore, it can be seen that nearly identical test
specimens in the same test program achieved considerably different
ρi =ρtotal is increased to 0.45%, but this is an 83% increase in the strength for some tests. This also supports potential anomalies in
quantity of steel. This result shows that adding internal reinforce- the test setup or procedure.
ment is a less efficient mechanism for increasing flexural resistance
than normal CFT members. To some extent this is compensated for
by the radius of the internal reinforcement. Increasing the rbm =rm Axial Load with No Lateral Loading
ratio increases the moment capacity because of the relatively larger A series of 48 models were analyzed under pure axial load. Again,
moment arm for the reinforcing steel. the predicted resistance values were compared to expressions avail-
Figs. 7(b and c) compare the moment capacity from finite- able in current design codes (AISC 2010; CEN 2004). Four differ-
element analysis to moment capacity computed by the AISC ent L=D ratios were considered as follows: 8, 12, 16, and 20. An
(2010) and Eurocode 4 (2004) provisions as a function of the initial imperfection with maximum amplitude of L=1,500, which is
ρi =ρtotal and D=t ratios, respectively. These analytical results show the value used to construct the AISC buckling curve, was used.
that current design codes provide reasonably conservative values The maximum load obtained from the analyses defines the
for moment capacity regardless of D=t ratio and ρi =ρtotal . On aver- bucking load. Fig. 8 shows the normalized axial load lateral drift
age, the AISC and Eurocode 4 provisions underestimate the resis- relationship curves for four models; the axial load was normal-
tance predicted by finite-element analysis by 9.1 and 8.5%, ized to the squash load, Po . For slender column models (for exam-
respectively. ple, L=D ¼ 16 and 20 as shown in Fig. 8), the axial strength

© ASCE 04013012-7 J. Struct. Eng.

J. Struct. Eng. 2013.139.


Mo,FEM / Mo,FEM without internal rebar
1.75 1.75 Mo,FEM /Mo,AISC
rbm/rm=0.8
rbm/rm=0.933 Mo,FEM /Mo,EC4
1.50

Mo,FEM /Mo,Predicted
1.50

1.25 1.25

1.00 1.00

0.75 0.75
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
/ (b) /
(a) i total
i total
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

1.75 Mo,FEM /Mo,AISC 1.75 Mo,Test /Mo,AISC


Mo,FEM /Mo,EC4 Mo,Test /Mo,EC4
1.50 1.50
Mo,FEM /Mo,Predicted

Mo,Test /Mo,Predicted
1.25 1.25

1.00 1.00
Test results (CFT without rebar)
0.75 0.75
40 60 80 100 120 40 60 80 100 120
(c) D/t ratio (d) D/t ratio

Fig. 7. Results of parametric study of CFTs under bending: (a) variation of moment capacity with ρi =ρtotal ; (b) M o;FEM =Mo;predicted versus ρi =ρtotal ;
(c) Mo;FEM =M o;predicted versus D=t ratio; (d) M o;Test =M o;predicted versus D=t ratio (test results)

1.2 and Roeder et al. (2010)] result in different values of λ, but the
D/t=60, and i=2% impact on the predicted buckling capacity is small. The AISC
1.0 buckling curve provides good agreement with buckling strength
L/D=8 predicted by finite-element analysis for all D=t ratios and ρi for
0.8 both EIeff models (AISC 2010; Roeder et al. 2010). With the
P/Po,AISC

L/D=12 Eurocode 4 (CEN 2004) buckling curves, an imperfection factor,


0.6
α, of 0.21 provides a better estimation of buckling strengths for the
L/D=16 CFT models than buckling curve b, as shown in Fig. 9(c). Local
0.4
buckling of the steel tube did not occur in these analyses, and there
L/D=20 were no strength reductions due to local buckling. Bradford et al.
0.2
(2002) and Hu et al. (2003) noted that local buckling of circular
0.0 CFTs under axial load is unlikely to occur because the concrete
0.0 0.3 0.6 0.9 1.2 core and steel tube are in contact around the perimeter and the con-
Drift (%) fining pressure is applied to the concrete in all the radial directions.
The simulated and code-predicted buckling capacities are com-
Fig. 8. Normalized axial load drift relationship for various L=D ratios pared in Fig. 10, which shows the normalized buckling load ratio
Pcr;FEM =Pcr;AISC or EC4 , where Pcr;AISC and Pcr;EC4 are the buckling
capacities from AISC (2010) and Eurocode 4 (CEN 2004), as a
continuously increased after the buckling without strength degra- function of the buckling parameter λ. The comparison suggests that
dation. For these more slender columns, the buckling load was the design equations provide a conservative estimate of the buck-
obtained by using the Southwell plot procedure (Mandal and ling capacity, and the conservatism is generally larger with increas-
Calladine 2002). Considerable prebuckling deformation was ob- ing λ for all design methods. The AISC buckling curve with EIeff
served for all models as shown in Fig. 8 and the magnitude of the from the AISC provision shows the best estimation of the buck-
deformation increased with increasing L=D ratio. These prebuck- ling strength. The average discrepancies were 9.5, 13, and 16.9%
ling deformations result in a secondary bending moment due to for AISC provisions when computed with EIeff expressions from
P-Δ effect. AISC (2010), Roeder et al. (2010), and Eurocode 4 (CEN 2004)
Fig. 9 compares the normalized buckling load Pcr;FEM = provisions, respectively.
Po;AISC or EC4 , where Pcr;FEM is the buckling strength obtained from Fig. 11 compares the predicted buckling strength with previous
finite-element analysis as a function of the buckling parameter λ. CFT test results without internal reinforcement. A total of 98 mea-
Different EIeff stiffness models as given by Eqs. (4) and (5) were sured buckling data points were used; all data are available on the
used to evaluate the AISC buckling provisions. Figs. 9(a and b) Goode CFT database (Goode and Lam 2008). The test specimens
show that the two different EIeff expressions [from AISC (2010) were selected to be of reasonable scale (diameter greater than

© ASCE 04013012-8 J. Struct. Eng.

J. Struct. Eng. 2013.139.


FEM, D/t=60, i=0%, FEM, D/t=100, i=0%
FEM, D/t=60 i=2%, rbm/rm=0.8, FEM, D/t=100 i=2%, rbm/rm=0.8

Eurocode 4 buckling curve a


1.0 1.0 1.0

0.8 0.8 0.8

Pcr,FEM /Po,AISC
Pcr,FEM /Po,AISC

Pcr,FEM /Po,EC4
0.6 0.6 0.6
Eurocode 4
0.4 0.4 0.4 buckling curve b
AISC buckling
AISC buckling curve
curve with EIeff
0.2 0.2 0.2
from Roeder et al. (2010)
0.0 0.0 0.0
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
(a) (b) (c) Buckling parameter
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

Buckling parameter Buckling parameter

Fig. 9. Variation of buckling load with λ: (a) Pcr =Po versus λ (AISC buckling curve); (b) Pcr =Po versus λ [AISC buckling curve with EIeff from
Roeder et al. (2010)]; (c) Pcr =Po versus λ (buckling curve from Eurocode 4)

1.50 1.50 1.50


1.25 1.25 1.25
Pcr,FEM /Pcr,AISC
Pcr,FEM /Pcr,AISC

Pcr,FEM /Pcr,EC4
1.00 1.00 1.00
0.75 0.75 0.75
0.50 0.50 AISC buckling curve with 0.50
AISC buckling curve Buckling curve a from
0.25 with EIeff from AISC 0.25 EIeff from Roeder et al. (2010) 0.25 Eurocode 4
0.00 0.00 0.00
0.5 1.0 1.5 2.0 0.5 1.0 1.5 2.0 0.5 1.0 1.5 2.0
(a) Buckling parameter (b) Buckling parameter (c) Buckling parameter

Fig. 10. Comparison of predicted buckling load: (a) comparison of analysis results with AISC buckling curve; (b) comparison of analysis results with
AISC buckling curve [EIeff from Roeder et al. (2010)]; (c) comparison of analysis results with buckling curve a from Eurocode 4 (CEN 2004)

2.5 diameters with small D=t ratios. Thus, their buckling behavior is
likely controlled by the steel section, and it is likely that the end
loading of the columns was imperfect, resulting in a coupled
2.0
global-local instability mode. Furthermore, nearly identical test
Pcr,Test /Pcr,Predicted

specimens in the same test program achieved significantly larger


1.5 (an average of 20%) compressive resistance. This suggests poten-
tial anomalies in the test setup or procedure for these three tests.
1.0

CFTs under Combined Loading


AISC buckling curve with EIeff from AISC
0.5
AISC buckling curve with EIeff from Roeder et al. (2010) A series of parametric analyses (a total of 113 models) were per-
Buckling curve a from Eurocode 4 formed to evaluate the strength of the CFT under general combined
0.0 loading. The analytical results were compared with the P-M inter-
0.0 0.5 1.0 1.5 2.0 2.5
Buckling parameter action curve from AISC (2010) and Eurocode 4 (CEN 2004).
Fig. 12 shows a P-M interaction curve derived from the plastic
Fig. 11. Comparison of test results with predicted buckling load stress distribution method and the analytical prediction of the
strength of CFT with a D=t of 60, L=D of 8, and ρi ¼ 0%. The
triangles denote the moment capacity when P-Δ moments are
included and the circles indicate the nominal moment capacity ex-
100 mm) and a resulting buckling parameter λ larger than 0.4. cluding P-Δ moments. When P-Δ moments are considered, the
Fig. 11 shows that the buckling strength predicted by current design P-M interaction curve from the plastic stress distribution method
provisions is generally conservative compared with the test results is the same as the curve that includes instability up to the buckling
and the predictions are more conservative with increasing λ. The load (Pcr;AISC =Po;AISC ¼ 0.8 for the model in Fig. 12). Conversely,
average discrepancies were 23.5, 24.2, and 26.7% for AISC pro- when the axial load exceeds buckling capacity, the plastic stress
visions, AISC buckling curve with EIeff from Roeder et al. (2010), distribution method overestimates the moment capacity of the
and Eurocode 4 (CEN 2004) provisions, respectively. A limited CFT. In this study, P-Δ effects were included in the moment
number of tests fall below the design provisions for columns with calculations for all comparisons.
λ values in the range between 0.5 and 1.0 (three test data points Fig. 13 compares P-M interaction curves based on AISC (2010)
out of a total of 98 points were below 95% of predicted buckling and Eurocode 4 (CEN 2004) provisions with finite-element analy-
load). These three test specimens in question have relatively small sis results for a CFT column having a D=t ratio of 60 and rbm =rm

© ASCE 04013012-9 J. Struct. Eng.

J. Struct. Eng. 2013.139.


PSDM
1.2 desirable, it has negative consequences with seismic design. Seis-
FEM results without P- effect
FEM results with P- effect mic design requires consideration of the expected capacity of
1.0 ductile elements because seismic performance does not depend
on the capacity of a single structural element. For example, if
0.8 Pcr,AISC=0.8Po,AISC
the conservatism of strength of a CFT column is large, the connec-
P/Po,AISC

0.6 tions and adjacent member such as footing or cap beam should
be conservatively designed to avoid premature failure of these
0.4 components in an extreme loading case. Thus, it is necessary to
accurately predict the strength of the CFT column. Furthermore,
0.2 most practical columns have relatively small axial load ratios
D/t=60, L/D=8, i=0%
0.0
and the AISC and Eurocode 4 interaction curves considerably
0.0 0.5 1.0 1.5 underestimate the moment capacity of CFTs under these condi-
M/Mo,AISC tions. This results in increased uncertainty as to the expected mo-
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

ment capacity of the member and requires greater conservatism in


Fig. 12. P-Δ effect on the strength of CFTs under combined loading the seismic design. Taken as a whole, these results suggest a need to
improve the P-M interaction curve to provide a more economical
and rational design.
of 0.8. Both reinforced and unreinforced CFT sections were con-
sidered. The figure shows that the plastic-stress distribution method
provides accurate and conservative predictions of the strength Alternative P-M Interaction Curve for Circular CFTs
of the CFT when the axial load is low, but as noted previously,
this method significantly overestimates the strength with high An alternative P-M interaction curve is proposed to improve the
axial load and slenderness ratio, because global buckling is not accuracy of the predicted resistance of CFT members under com-
considered in the P-M interaction curves from the plastic-stress bined load. The proposed curve is shown in Fig. 14 and is specially
distribution method. designed to take the benefits from both the plastic-stress distribu-
If the axial load is limited to the buckling resistance, the plastic- tion method and current design codes. The curve is defined as
stress distribution method may slightly overestimate the moment follows:
capacity as the compressive load approaches the buckling, but • The P-M interaction curve from the PSDM forms the basis of
the overestimate is modest. The interaction curve suggested by the curve.
AISC (2010) and Eurocode 4 (CEN 2004) interaction curves were • Points A and B are the pure axial and pure flexural capacity cor-
similar to each other and give quite conservative estimation of the responding to the PSDM P-M interaction curve (no considera-
strength of the CFT for all analyzed cases. While conservatism is tion of global buckling effects).

1.2 D/t=60, L/D=8, i=0% 1.2 D/t=60, L/D=12, i=0%


PSDM
PSDM 1.0
1.0 AISC
AISC
Eurocode 4
Eurocode 4 0.8
0.8 Proposed
Proposed
FEM
P/Po,AISC

P/Po,AISC

FEM
0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0.0 0.5 1.0 1.5 0.0 0.5 1.0 1.5
(a) M/Mo,AISC (b) M/Mo,AISC

1.2 D/t=60, L/D=8, =2%, r /r =0.8 1.2 D/t=60, L/D=12, =2%, r /r =0.8
i bm m i bm m

1.0 PSDM
1.0 PSDM
AISC AISC
0.8 0.8 eurocode
P/Po,AISC

Eurocode 4
P/Po,AISC

Proposed proposed
0.6 FEM 0.6 FEM

0.4 0.4

0.2 0.2
0.0 0.0
0.0 0.5 1.0 1.5 0.0 0.5 1.0 1.5
(c) M/Mo,AISC (d) M/Mo,AISC

Fig. 13. Comparison of P-M interaction curves with analysis results: (a) D=t ¼ 60, L=D ¼ 8, ρi ¼ 0%; (b) D=t ¼ 60, L=D ¼ 12, ρi ¼ 0%;
(c) D=t ¼ 60, L=D ¼ 8, ρi ¼ 2%; (d) D=t ¼ 60, L=D ¼ 12, ρi ¼ 2%

© ASCE 04013012-10 J. Struct. Eng.

J. Struct. Eng. 2013.139.


1.2 D/t=60, L/D=8, i=0% 3.0
A
1.0 2.5
A' A'' Interaction curve from PSDM
0.8

Mu,Test /Mu,Predicted
P/Po,AISC 2.0
C
0.6
1.5
0.4 Proposed interaction curve
C'
including global buckling effect 1.0
D AISC (AVG.=1.59, STD.=0.39)
0.2
0.5 Eurocode 4 (AVG.=1.58, STD.=0.39)
B Proposed curve (AVG.=1.29, STD.=0.17)
0.0
0.0 0.5 1.0 1.5 0.0
M/Mo,AISC 0.0 0.1 0.2 0.3 0.4 0.5
P/Pcr,AISC
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Schematic view of alternative P-M interaction curve


Fig. 16. Comparison of test results with predicted strength

• Point C lies on the PSDM P-M interaction curve with the same
moment capacity as point B. The different between points B and Fig. 16 compares the predications using the proposed curve with
C is that point B does not have an axial load and point C does. previous test results (Elremaily and Azizinamini 2001; Bishop
• Points A 0 and C 0 are obtained by from points A and C by multi- 2009; Zhang et al. 2009; Park et al. 1982). A total of 34 test results
plying the axial load associated with points A and C by the ratio were compared. The D=t ratio of test specimens varied from 34.5 to
Pcr =Po;AISC . (This is similar to the AISC approach.) These 111.9, and axial load ratio P=Pcr had a range from 0.06 to 0.44.
points do not fall on the PSDM P-M interaction curve. [Only the tests of Park et al. (1982) had internal reinforcement.]
• Point D is located on the PSDM interaction curve. The point The proposed design curve provides the conservative predictions
is determined as PD ¼ 0.5PC 0 , where PD is the axial load at of all of the specimens. On average, the results provided moment
point D. capacities that were 29, 59, and 58% larger than the moment
• Finally, the maximum axial load, PA 0 , is used to determine the capacity predicted by the proposed model and AISC (2010) and
upper limit of the curve. The intersection of the PSDM P-M Eurocode 4 (CEN 2004) provisions, respectively. This increased
interaction curve and the horizontal line at PA 0 is used. capacity is expected; work by Roeder et al. (2010) shows that
• The alternative P-M interaction curve can be constructed by the PSDM method results in larger capacities than predicted,
connecting points A 0 , A 0 0 , D, and B. 24% on average. Taken as a whole, this thorough comparison in-
The proposed curve is also shown in Fig. 13. The proposed dicates that the proposal is accurate, conservative, and appropriate
curve compares well with the simulated results. Fig. 15 compares for design.
the ratio of moment predicted by the finite-element analysis (total
of 113 models) to the moment capacity predicted using AISC
(2010), Eurocode 4 provisions (CEN 2004), and proposed P-M in- Summary and Conclusions
teraction curve. The ratio is plotted as a function of the applied axial
load. The proposed interaction curve provides more accurate esti- In this study, the strength and behavior of CFTs with and without
mate and less scatter than the provisions, as indicated by the mean internal reinforcement under general combined loading was stud-
and standard deviation values presented in the figure. The scatter ied. The research approach was as follows. A finite-element model
increases as the axial load P approaches the critical load Pcr . Larger capable of simulating generalized conditions, including the impact
values of P beyond balance results in smaller moment capacities, of confinement, slip, and axial load, was developed. The model
therefore inaccuracies may be amplified. The proposed curve pro- was verified using previous tests for CFTs under various loading
vides more accurate flexural strength predictions at all levels of ax- conditions and the accuracy of the finite-element models was veri-
ial loads. Fig. 15 also shows a magnified portion of the curve at fied. Parametric studies were conducted using the verified model.
more common levels of axial load ratios; the proposed curve is both The study parameters included amount of internal reinforcement,
accurate and conservative. ratio of the tube to the internal reinforcement, axial load ratio,

AISC (AVG.=1.76 , STD.=1.00) 1.75 AISC (AVG.=1.18 , STD.=0.07)


6
EC4 (AVG.=2.10 , STD.=2.07) Eurocode 4 (AVG.=1.20 , STD.=0.09)
Proposed curve Proposed curve
5 (AVG.=1.07 , STD.=0.05)
Mu,FEM/Mu,Predicted

1.50
Mu,FEM/Mu,Predicted

(AVG.=1.06 , STD.=0.06)
4

3 1.25

2
1.00
1

0 0.75
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6
P/Pcr,AISC P/Pcr,AISC

Fig. 15. Comparison of analysis results with predicted strength

© ASCE 04013012-11 J. Struct. Eng.

J. Struct. Eng. 2013.139.


slenderness ratio, and D=t ratio. These parameters were evaluated Elremaily, A., and Azizinamini, A. (2001). “Design provisions for connec-
for all combined loading conditions. tions between steel beams and concrete filled tube columns.” J. Constr.
The results were used to both evaluate existing P-M interaction Steel Res., 57(9), 971–995.
curve design models and a proposed model. The AISC (2010) and European Committee for Standardization (CEN). (2004). “Design of
composite steel and concrete structures, Part 1.1 General rules and rules
Eurocode 4 (CEN 2004) provisions provide P-M interaction curves
for buildings.” EN 1994-1-1, Eurocode 4, Brussels, Belgium.
including consideration of buckling effects, but these curves con- Goode, C. D., and Lam, D. (2008). “Concrete—filled tube columns—tests
siderably underestimate the bending resistance of CFTs for the low compared with Eurocode 4.” Proc., Engineering Foundation Conf. on
axial loads. This is the most practical range for column design, and Composite Construction in Steel and Concrete, ASCE, Reston, VA,
an alternate P-M interaction curve was proposed. 317–325.
Under pure bending, the analysis showed that the AISC (2010) Han, L. H., Lu, H., Yao, G. H., and Liao, F. Y. (2006). “Further study on the
and Eurocode 4 (CEN 2004) provisions provide reasonable esti- flexural behaviour of concrete-filled steel tubes.” J. Constr. Steel Res.,
mates of flexural capacity of the CFTs for all values of D=t and 62(6), 554–565.
internal reinforcing bar ratio. Under axial compression, the AISC Han, L.-H., and Yan, S.-Z. (2000). “Experimental studies on the strength
with high slenderness ratio concrete filled steel tubular columns.” Proc.,
Downloaded from ascelibrary.org by East Carolina University on 06/15/14. Copyright ASCE. For personal use only; all rights reserved.

buckling curve agrees well with the results of simulated and test
6th ASCCS Conf., ASCCS, West Yorkshire, UK, 419–425.
results. Under general combined loading, the P-M interaction curve
Hu, H. T., Huang, C. S., Wu, M. H., and Wu, Y. M. (2003). “Nonlinear
from the plastic-stress distribution method considerably overesti- analysis of axial loaded concrete-filled tube columns with confinement
mates the strength when the slenderness and axial load ratios effect.” J. Struct. Eng., 129(10), 1322–1329.
are high. This is expected because the plastic-stress distribution Lee, J., and Fenves, G. L. (1998). “Plastic-damage model for cyclic loading
method neglects the effect of global buckling. of concrete structures.” J. Eng. Mech., 124(8), 892–900.
The alternative P-M interaction curve reduced the significant Leon, R. T., and Hu, J. W. (2008). “Design of innovative SMA PR con-
conservatism noted in current AISC (2010) and Eurocode 4 nections between steel beams and composite columns.” Proc., 6th
(CEN 2004) methods and provided reduced scatter. The proposed Int. Workshop on Connections in Steel Structures, Connections VI,
curve employs aspects of both the plastic-stress distribution method AISC, Chicago, IL, 513–524.
and the current design codes. The proposed curve was successfully Leon, R. T., Kim, D. K., and Hajjar, J. F. (2007). “Limit state response of
composite columns and beam-columns part 1: Formulation of design
verified by comparing with results of finite-element analysis and
provisions for the 2005 AISC specification.” Eng. J., 44(4), 341–358.
those of previous tests, and it was shown to provide more accurate Lubliner, J., Oliver, J., Oller, S., and Oñate, E. (1989). “A plastic-damage
predictions of combined resistance than current design methods. model for concrete.” Int. J. Solids Struct., 25(3), 299–326.
Mandal, P., and Calladine, C. R. (2002). “Lateral-torsional buckling of
beams and the Southwell plot.” Int. J. Mech. Sci., 44(12), 2557–2571.
Acknowledgments Marson, J., and Bruneau, M. (2004). “Cyclic testing of concrete-filled cir-
cular steel bridge piers having encased fixed-base detail.” J. Bridge
This paper was completed partially with funding provided by the Eng., 9(1), 14–23.
Washington State Department of Transportation (WSDOT) through Matsui, C., Tsuda, K., and Ishibashi, Y. (1995). “Slender concrete filled
the project entitled “Design of Bridge Foundations with Steel Cas- steel tubular columns under combined compression and bending.”
Proc., 4th Pacific Structural Steel Conf., Steel-Concrete Composite
ings.” Mr. Bijan Khaleghi is the WSDOT Bridge Design Engineer
Structures, Vol. 3, Pergamon, Oxford, UK, 29–36.
and the coordinator of this project. The advice and financial support Moon, J., Lehman, D. E., Roeder, C. W., and Lee, H.-E. (2012). “Analytical
of the WSDOT is gratefully acknowledged. modeling of bending of circular concrete-filled steel tubes.” Eng.
Struct., 42, 349–361.
Park, R. J. T., Priestley, M. J. N., and Walpole, W. R. (1982). “The seismic
References performance of steel encased reinforced concrete bridge piles.”
Research Rep. 82-12, Civil Engineering Dept., Univ. of Canterbury,
AASHTO. (2009). AASHTO LRFD bridge design specification, 4th Ed., Christchurch, New Zealand.
Washington, DC. Prion, H. G. L., and Boehme, J. (1994). “Beam-column behaviour of
ABAQUS [Computer software]. Dassault Systemes Simulia Corp., steel tubes filled with high strength concrete.” Can. J. Civ. Eng.,
Providence, RI. 21(2), 207–218.
American Concrete Institute (ACI). (2008). Building code requirements Roeder, C. W., Lehman, D. E., and Bishop, E. (2010). “Strength and
for structural concrete and commentary, Farmington Hills, MI. stiffness of circular concrete filled tubes.” J. Struct. Eng., 136(12),
American Institute of Steel Construction (AISC). (2010). Specifications 1545–1553.
Roik, K., and Bergmann, R. (1992). “Composite column.” Constructional
for structural steel buildings, Chicago.
steel design: An international guide, P. J. Dowling, J. E. Harding, and
Bishop, E. S. (2009). “Evaluation of the flexural resistance and stiffness
R. Bjorhovde, eds., Elsevier Applied Science, London.
models for circular concrete-filled steel tube members subjected to com-
Saenz, L. P. (1964). “Discussion of ‘Equation for the stress-strain curve
bined axial-flexural loading.” M.S. thesis, Univ. of Washington, Seattle.
of concrete’ by P. Desayi, and S. Krishnan.” J. Am. Concr. Inst., 61,
Bradford, M. A., Loh, H. Y., and Uy, B. (2002). “Slenderness limits for 1229–1235.
filled circular steel tubes.” J. Constr. Steel Res., 58(2), 243–252. Thody, R. (2006). “Experimental investigation of the flexural properties
Bruneau, M., and Marson, J. (2004). “Seismic design of concrete-filled of high-strength concrete-filled steel tubes.” M.S. thesis, Univ. of
circular steel bridge piers.” J. Bridge Eng., 9(1), 24–34. Washington, Seattle.
Choi, Y.-H., Foutch, D. A., and LaFave, J. M. (2006). “New approach Varma, A. H., Ricles, J. M., Sause, R., and Lu, L. W. (2002). “Experimental
to AISC P-M interaction curve for square concrete filled tube (CFT) behavior of high strength square concrete-filled steel tube beam-
beam-columns.” Eng. Struct., 28(11), 1586–1598. columns.” J. Struct. Eng., 128(3), 309–318.
Elchalakani, M., Zhao, X. L., and Grzebieta, R. H. (2001). “Concrete-filled Zhang, G. W., Xiao, Y., and Kunnath, S. (2009). “Low-cycle fatigue dam-
circular steel tubes subjected to pure bending.” J. Constr. Steel Res., age of circular concrete-filled tube columns.” ACI Struct. J., 106(2),
57(11), 1141–1168. 151–159.

© ASCE 04013012-12 J. Struct. Eng.

J. Struct. Eng. 2013.139.

Potrebbero piacerti anche