You are on page 1of 18

Structures 26 (2020) 870–887

Contents lists available at ScienceDirect

Structures
journal homepage: www.elsevier.com/locate/structures

Bending strength and design methods of the 6082-T6 aluminum alloy beams T
with circular hollow sections
Yuanzheng Zhao, Ximei Zhai

School of Civil Engineering, Harbin Institute of Technology, Harbin 150090, PR China

ARTICLE INFO ABSTRACT

Keywords: Circular hollow section (CHS) specimens are widely employed in long-span space structures, truss structures, and
6082-T6 aluminum alloy bridge structures; however, only a few experimental and numerical investigations on CHS aluminum alloy beams
Circle hollow section have been conducted in China. This paper presents a detailed experimental investigation on CHS beams made of
Bending strength 6082-T6 aluminum alloys. The effects of length and section slenderness on the ultimate strength and buckling
Reliability analysis
behaviour were investigated. Seventeen extruded CHS specimens were subjected to four-point bending tests,
Finite element analysis
Direct strength method
with the diameter-to-thickness ratio (D/t) ranging from 16.4 to 29.1. The failure modes observed included
Continuous strength method overall buckling and the coupling of local and overall buckling. The initial geometric imperfections were
measured prior to the bending tests. Subsequently, fine finite-element (FE) models were developed using the
nonlinear analysis program ABAQUS and validated against the experimental results. An extensive parametric
study involving 600 numerical results was performed to evaluate the effects of D/t, the section dimensions, and
the slenderness ratio on the mechanical responses and bending strength of the CHS beams. The test results,
together with the obtained FE analysis results, were utilised to assess the accuracy of the bending-strength
provisions in the current design codes GB 50429 (China), Eurocode 9 (Europe), and AA-2015 (America), as well
as optional design approaches, i.e. the direct strength method (DSM) and continuous strength method (CSM).
The results indicated that all five design provisions provided conservative predictions of the ultimate moment
capacities; the predicted results of DSM and CSM were more accurate than the three design specifications. The
reliability levels of the bending-strength provisions were confirmed and compared using statistical parameters
from the corresponding specifications.

1. Introduction including 26 deep rectangular beams and 10 I-section beams made of


17 ST and 27 ST aluminum alloys, via experimental and analytical
The applications of aluminum alloys are growing rapidly in the studies. The lateral buckling of deep beams agreed with the classical
construction industry, owing to their favourable corrosion resistance, theory proposed by Prandtl et al., and the authors also established an
high strength-to-weight ratios, excellent appearance, and ease of approximate formula for calculating the critical stress of the beams.
maintenance [1–4]. Aluminum alloys are mainly adopted as load- Clark et al. [9] investigated the elastic and inelastic lateral buckling of
bearing materials in space structures and bridge structures; however, I-beams subjected to unequal end moments and proposed an approx-
their relatively low Young’s modulus (approximately 1/3 of that of imate design formula for the lateral strength. Mazzolani et al. [10–12]
carbon steel) can have adverse effects on the buckling deformations and investigated the rotation capacity, bearing capacity, and deformation
bending strength of the aluminum alloy beams [5]. Additionally, for behaviours of aluminum alloy beams, and a new equation for calcu-
specimens with large diameter-to-thickness ratios (D/t) and small lating the lateral-torsional stability coefficient was introduced, which
slenderness ratios ( ), local buckling, together with overall buckling, differed from the traditional Perry–Robertson formula. Scholars from
can significantly affect the ultimate moment capacities and should not the Imperial College London and University of Hong Kong (Gardner,
be ignored. Zhu, Young, and Su) [13–16] performed systematic tests and numerical
American and European researchers have conducted experimental analyses of aluminum alloy beams and investigated the design effi-
studies on aluminum alloy beams since the 1930s. Dumont and Hill ciency of the direct strength method (DSM) and continuous strength
[6–8] examined the lateral instability of aluminum alloy beams, method (CSM). Moen et al. [17–19] performed detailed experimental


Corresponding author.
E-mail address: xmzhai@hit.edu.cn (X. Zhai).

https://doi.org/10.1016/j.istruc.2020.05.007
Received 7 January 2020; Received in revised form 19 April 2020; Accepted 5 May 2020
Available online 20 May 2020
2352-0124/ © 2020 Institution of Structural Engineers. Published by Elsevier Ltd. All rights reserved.
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Nomenclature Mel elastic moment capacity (Mel = Wel f0.2 )


Mpl plastic moment capacity (Mpl = Wpl f0.2 )
COV coefficient of variation Nu,test ultimate vertical load obtained from the experiments
D nominal outer diameter of CHS Nu,FEA ultimate vertical load obtained from the simulations
D0 measured outer diameter of CHS Ne ultimate elastic load (Ne = 8Wel f0.2 Le )
D/t diameter-to-thickness ratio of CHS specimens n hardening exponent
Dk dead load O overall flexural buckling
fu ultimate tensile stress t nominal wall thickness of specimens
f0.2 0.2% proof stress (nominal yield strength) te effective wall thickness of specimens
E0 initial Young’s modulus t0 measured wall thickness of specimens
i radius of gyration of the cross-section Wel elastic modulus of gross section
L local flexural buckling Wex effective elastic modulus of gross section
La actual length of the specimens Wpl plastic modulus of gross section
Le effective length of the columns (Le = La-200 mm) slenderness ratio ( = Le i )
Lk live load v0 amplitude of the initial imperfections
M0.1 experiment moment with residual deflection of 0.1% Le load ratio ( = Lk D k )
MAA ultimate moment predicted by AA-2015 σ stress
MCSM ultimate moment predicted by CSM ε strain
MDSM ultimate moment predicted by DSM ξ slenderness parameter of the cross-section
MEU ultimate moment predicted by Eurocode 9 γx cross-section plastic adaption coefficient
MGB ultimate moment predicted by GB 50429 resistance factor
Mu,test ultimate moment of the experimental beams β reliability index
Mu,FEA ultimate moment of the numerical beams [β] target value of reliability index

and numerical investigations of aluminum alloy beams to investigate strength and mechanical behaviours of aluminum alloy CHS beams and
the rotation capacity and ultimate bending strength, and the classifi- to assess the suitability and reliability levels of several common design
cation of the cross sections in Eurocode 9 was discussed. previsions for the 6082-T6 aluminum alloy CHS beams. First, through a
Research on aluminum alloy beams started rather late in China. Wu detailed test of 17 extruded specimens, the buckling behaviours, ulti-
[20] and Guo [21–22] conducted a series of simulated and experi- mate bending strength, and deformation capacity of 6082-T6 aluminum
mental research on beams with I-sections, T-sections, and channel alloy CHS members subjected to four-point bending were investigated.
sections and proposed design formulas for the flexural-torsional buck- Second, finite-element (FE) models were developed using the FE ana-
ling coefficient. The lateral-torsional buckling resistance of I-beams lysis (FEA) software ABAQUS and verified according to the test results.
subjected to pure bending was investigated by Wang and Yuan [23]; the Then, they were adopted for a parametric study, where the effects of D/
design rules in Eurocode 9 [24] and a modified design method based on t, the section dimensions, and the loading positions on the structural
the Chinese code for steel structures were evaluated according to the response and bending strength of CHS beams were investigated. Third,
test results. Chen and Feng [25,26] performed an experimental study on the experimental and numerical results were compared with the design
CFRP (carbon fibres reinforced polymers) strengthened concrete-filled strengths predicted using the Chinese code (GB 50429) [27], European
aluminum alloy CHS beams and investigated the failure modes and code (Eurocode 9) [24], America code (AA-2015) [28], DSM [29], and
ultimate strengths of the composite specimens. However, experimental CSM [30] for evaluating their accuracy. The reliability levels of the five
data and advanced design methods for the bending strength of alu- design provisions were assessed according to the test and simulated
minum alloy beams in China are lacking. results.
The objective of the present study was to examine the bending

Table 1
Measured dimensions and key results from the bending tests.
Specimen D (mm) T (mm) Le (mm) v0/Le (‰) M0.1 (kN·m) Mu,test (kN·m) Fu,test (kN·m) Failure modes

CHS-89-4-1800A 88.62 3.80 1800 59.97 0.558 6.17 8.24 36.62 O +L


CHS-89-4-1800B 88.63 3.88 1800 60.01 0.147 6.24 8.06 35.82 O +L
CHS-89-4-1800C 88.83 3.90 1800 59.88 0.529 6.06 8.47 37.64 O +L
CHS-89-4-2200A 88.41 3.92 2200 73.56 0.694 6.05 7.66 27.85 O
CHS-89-5-1800A 88.89 4.97 1800 60.56 0.860 9.16 11.21 49.82 O
CHS-89-5-1800B 89.12 5.08 1800 60.47 0.426 9.01 10.67 47.42 O
CHS-89-5-2200A 88.82 4.99 2200 74.09 0.948 8.50 10.63 38.65 O
CHS-89-5-2200B 88.94 4.99 2200 73.99 0.632 9.05 11.09 40.33 O
CHS-114-4-1800A 114.22 3.93 1800 46.13 0.874 9.70 16.00 71.11 O +L
CHS-114-4-2200A 114.42 4.03 2200 56.33 1.241 10.38 14.17 51.53 O +L
CHS-114-4-2200B 113.96 4.06 2200 56.58 0.224 11.42 14.50 52.73 O +L
CHS-114-5.5-1800A 113.84 5.55 1800 46.95 0.882 15.32 18.93 84.13 O
CHS-114-5.5-2200A 114.16 5.45 2200 57.17 0.801 15.77 19.24 69.96 O
CHS-114-7-1800A 113.99 6.90 1800 47.44 0.831 21.93 25.61 113.82 O
CHS-114-7-1800B 113.46 6.76 1800 47.62 0.905 20.16 24.81 110.27 O
CHS-114-7-2200A 113.61 6.90 2200 58.19 0.611 23.59 26.03 94.65 O
CHS-114-7-2200B 113.84 6.90 2200 58.06 1.321 21.46 24.83 90.29 O

871
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

2. Experimental study The dimensions of the rings are shown in Fig. 3. A concentrated vertical
load was applied to the tested specimens at the mid-spans via a hy-
2.1. Test specimens and initial geometric imperfections draulic jack. A spreader beam was placed under the jack, and the
concentrated load was transferred onto the CHS specimens at two
A total of 17 CHS specimens, which were extruded by a 6082-T6 points, as shown in Fig. 4. Each loading point was located at a distance
aluminum alloy, were tested under four-point bending. A pure bending of Le/4 (450 or 550 mm) from the centreline of the steel ring support,
region was obtained between the loading points; the main design and the length of the pure bending region was taken as Le/2 (900 or
parameters of the specimens were the diameter-to-thickness ratio D/t 1100 mm). The beam ends and loading points of the specimens were
and the slenderness ratio . Five section dimensions (89 mm × 4 mm, designed to rotate freely in the bending plane; thus, cylindrical rollers
89 mm × 5 mm, 114 mm × 4 mm, 114 mm × 5.5 mm, and were placed at the end supports and between the spreader beam and the
114 mm × 7 mm), were adopted for the tested CHS beams with D/t sleeves mounted on the specimens.
values ranging from 16.4 to 29.1. Two actual lengths (La) of 2400 mm The instruments consisted of a load cell, seven linear variable dif-
and 2000 mm were selected for each cross section, with varying from ferential transformers (LVDTs), and 12 strain gauges, as shown in
46.1 to 76.1. One or two repeated tests were performed on most of the Fig. 5(a). The load cell was sandwiched between the hydraulic jack and
tested beams to validate the experimental results. In this test, all the the counterforce beam to measure the applied vertical load, and the
specimens were simply supported, and the exerted lengths of both ends sensitivity coefficient of the cell was calibrated using the hydraulic
beyond the centrelines of the supports were taken as 100 mm; thus, the press before the bending tests. Three 150-mm LVDTs (DS-5, DS-6, and
effective lengths (Le) were reduced to 2200 and 1800 mm. Prior to the DS-7) were located under the tested specimens at the mid-span and the
bending tests, the actual dimensions of the cross sections were mea- two loading points to record the vertical displacements, and two 25-mm
sured for each beam, as shown in Table 1, and the specimen label LVDTs (DS-1/DS-2 or DS-3/DS-4) were positioned on the side surface of
contained the nominal dimensional details, i.e. the diameter D, thick- the rings at each end support to measure the rotation capacities of the
ness t, and effective length Le. The letters A, B, and C at the end of the specimen ends. Three groups of strain gauges were attached to the
specimen labels represent the repeated tests of the CHS beams. surfaces of the specimens along the length direction, with two groups
The initial imperfections of the metal specimens mainly included near the loading points and one group on the mid-span. The locations of
the residual stress and geometric imperfections, which can significantly the strain gauges are shown in Fig. 5(b). A data-acquisition system
affect their ultimate strengths. However, for aluminum alloy specimens, (TST3826E) was adopted for collecting the experimental data for the
the residual stress (which is less than 20 MPa) can be ignored on ac- applied loads, displacements, and strains in the bending tests.
count of the extrusion moulding technique [31,32]; thus, only the in-
itial geometric imperfections were measured in this study, including the
shape and the amplitude values, as shown in Fig. 1. A laser displace- 2.3. Test procedure
ment transducer with a precision of 0.001 mm was employed to mea-
sure the geometric imperfections along the four generatrix lines on the Before the application of the external load, a preliminary load,
specimens’ surfaces. These lines were separated from each other by 90°. which was approximately 20% of the ultimate elastic load Ne (defined
The imperfections were recorded at intervals of 50 mm along the spe- as the load that can produce the nominal ‘yield’ moment and
cimen length, and their maximum values were taken as the imperfec- Ne = 8Wel f0.2 Le , where Wel = D3 (1 4 ) 32
and = (D 2t ) D in
tion amplitudes (v0), which are listed in Table 1 for each specimen. The this study) was applied to the tested specimens to verify the effective-
typical initial geometric imperfection curves are shown in Fig. 2 (for ness of the test setup and data-acquisition system. After the preloading
CHS-89-5-2200B and CHS-114-5.5-1800A). The imperfection curves of was removed, the concentrated vertical load was applied to the
the CHS specimens can each be simplified as a half-sinusoid curve with spreader beam under displacement control at a rate of 2 mm/min. The
an amplitude less than 1/1000 Le. loading system would have been stopped if either of the following two
conditions was satisfied:
2.2. Experimental setup and instrument configurations
(1) local buckling occurred and the vertical load reduced to approxi-
Considering the cylindrical shape of the CHS specimens, the ends of mately 70% of the ultimate value;
the tested beams were not placed directly on the support rollers, to (2) there was no local buckling, and the vertical displacement kept
avoid local failure caused by concentrated loads. To achieve this, four increasing under a constant load.
precision-machined steel rings with internal diameters equal to the
external diameter of the CHS were closely tied to the specimens at the There were other practical conditions that would have stopped the
two loading points and end supports [33]. Each ring consisted of two loading, e.g. the distance between the specimen and the edge of the end
parts, which were connected by bolts; one part included a cylindrical supports, the maximum stroke (200 mm) of the jack, and the measuring
roller and a half-sleeve, and the other part was simply the half-sleeve. ranges ( ± 25 or ± 200 mm) of the LVDTs.

Fig. 1. Measurement of the initial geometric imperfections.

872
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 2. The initial geometric imperfection curves.

Fig. 3. View of the steel ring.

2.4. Test results

2.4.1. Failure modes


All the tested beams exhibited overall buckling, and some specimens
exhibited a coupling mode of overall and local buckling, as shown in
Table 1. The local buckling occurred in the compressive region of the
specimens, and no material fracture was observed in the tensile region,
owing to the excellent ductility of the aluminum alloys. Because four-
point bending was adopted in this study, the middle part (between the
two loading points under the spreader beam) was subjected to pure
bending, and the buckling modes of the tested beams varied with re-
spect to D/t. For the beams with large diameter-to-thickness ratios (D/
t ≥ 20), local buckling occurred in the post-loading stage, which pre-
vented the vertical load from increasing further. In contrast, for the
beams with smaller diameter-to-thickness ratios (D/t < 20), the
growth of the loading capacity remained fairly slow in the post-loading
stage, without local buckling. The beams were unloaded when the de-
flection of the mid-spans was too large for continued loading, and
overall buckling was regarded as their failure mode. The typical failure
modes of the tested specimens are shown in Fig. 6. After the unloading,
Fig. 4. Experimental setup for four-point bending tests.
approximately half of the deformation of the beams that failed via
overall buckling was recovered, and the residual deformation shapes
were similar to the half-sinusoidal curves; however, little recovery oc-
curred for the beams that suffered local buckling. Fig. 7 shows the

873
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 5. Schematic diagram of the four-point bending test set-up.

residual deformation for all 17 tested beams. The local buckling oc- In the initial stage of the loading process, the loads increased line-
curred near the loading points rather than the exact mid-spans. Because arly with an increase in the mid-span displacement or the end rotation,
it was impossible to apply the concentrated load exactly at the centre of which was caused by the elasticity of the whole cross section. The
the spreader beam, the steel ring locations had small errors. growth rates in the linear stage were higher for the specimens with
The ultimate concentrated loads (Fu,test) and bending moments (M0.1 smaller lengths and diameter-to-thickness ratios. When the extreme
and Mu,test) are presented in Table 1, together with the buckling modes fibre of the cross section reached the yielding stress, the stress of the
of the specimens. Generally, the ultimate load Fu,test decreased with an cross section transitioned into the elastic–plastic stage, the duration of
increase in the slenderness ratio . On average, the bending moment which depended on the length and diameter-to-thickness ratios of the
with residual deflection of 0.1% Le (M0.1) was approximately 80% of tested beams. The vertical loads continued to increase with the de-
the ultimate bending moment (Mu,test), which was mainly determined creasing growth rate and finally became constant with the deformation
by the section dimensions (D and t). However, for the specimens with increasing continually. A sudden decrease in the vertical load appeared
the same cross sections, the differences of the bending moments were when the tested beams underwent local buckling, and the deflection or
not significant with the variation of , which ranged from 46.1 to 76.1 the end rotation kept increasing, with higher growth rates than before.
in this study. For the specimens that failed via overall buckling, the vertical load
continued to increase after the elastic–plastic stage because of the
2.4.2. Load–deformation curves material nonlinearity of the aluminum alloys, and the specimens were
The recorded deformation of the tested beams included the deflec- unloaded when there was no significant variation in the applied loads.
tion at the mid-spans and the rotation capacity of the end supports, Generally, the deformation capacity, including the mid-span deflection
which are plotted with respect to the vertical loads in Figs. 8 and 9. The and the end rotation capacities, was inversely proportional to the ul-
end rotation of the tested specimens was calculated using the dis- timate load, except for the specimen that failed via local buckling (CHS-
placement measured by the LVDTs (DS-1/DS-2 or DS-3/DS-4), which 114-4-1800A).
were positioned at the end supports, as shown in Fig. 5. In Figs. 8 and 9,
the load–deformation curves of the repeatedly tested specimens, such as 2.4.3. Load–strain curves
CHS-89-5-1800A and CHS-89-5-1800B, almost overlap, and the largest Figs. 10 and 11 show the vertical load–strain curves of the tested
difference in the ultimate load Fu,test did not exceed 5%. CHS beams obtained at the mid-spans and loading points, respectively.

Fig. 6. Typical failure modes of bending test specimens.

874
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 7. Residual deflections of all bending specimens.

The variation curves are plotted in two separate figures (Fig. 10(a) and 3. Numerical simulations
(b), Fig. 11(a) and (b)) according to the effective length Le of the spe-
cimens. The load–strain curves of the repeatedly tested beams almost 3.1. FE modeling
overlapped, similar to the corresponding load–deformation curves in
Figs. 8 and 9. The FE models of the 6082-T6 aluminum alloy beams were devel-
In Figs. 10 and 11, the positive and negative values represent tensile oped in parallel with the experimental investigations using the FE
and compressive strains, respectively. The measurement results for software package ABAQUS [34] and the typical model for the CHS
strains M-1 and M-3, which were located at the top and bottom of the specimens is given in Fig. 12(a).
cross section, exhibited an obvious symmetry variation. The strain Q-4 The steel rings located at the load application points were simplified
at the side of the surface was far lower (by a factor of approximately 10) as an integrated sleeve with a reference node, which was placed above
than the strain Q-1 at the top. Thus, the specimens underwent no sig- the center of upper surface of the sleeve. The steel sleeve was modeled
nificant lateral buckling or torsion during the loading process, and the using the solid element C3D8R (Fig. 12(b)) and the concentrated point
deformation of the cross section complied with the assumption of a loads with the same amplitude values were applied at the reference
plane section. The load–strain curves indicated linear growth in the nodes (RF-3 and RF-4). The general contact was applied to model the
initial loading stage, and the growth rate increased with the D/t of the connection between the inner surface of the steel sleeves (master sur-
beams. After the cross section reached the elastic–plastic stage, the face) and the outer surface of the beam specimens (slave surface); hard
vertical loads kept increasing with the decreasing growth rate, while contact and friction penalty contact (with the friction coefficient of 0.1)
the strains increased at a higher rate than before. Before the uploading, were assigned to the general contact properties in the normal direction
there existed a ‘yielding’ stage in the load–strain curves, which re- and tangential direction, respectively.
presented the extent to which the plastic deformation of the cross The steel rings located at the end supports were simplified as two
sections could develop. Generally, the length of the ‘yielding’ stage was reference nodes (RF-1 and RF-2) and they were connected to the end
inversely proportional to the ultimate load; conversely, the growth rate surface of the beam model by the surface-based “tie” constraint. The
of the vertical load was proportional to the ultimate load. The strain ends of the FE models were designed to be simply supported in ac-
values for the long specimens (Le = 2200 mm) were significantly lower cordance with the tested beams. Therefore, the reference nodes at each
than those for specimens that had the same cross section but were end support were free to rotate about the x–x axis and translate in the
shorter (Le = 1800 mm). longitudinal z–z axis while the other degrees of freedom were restrained
(U1 = U2 = 0; UR2 = UR3 = 0). On the other hand, the reference
nodes at the loading points were restrained against all degrees of
freedom except for the translation in the vertical y–y axis

Fig. 8. Vertical load-deflection curves at the mid-spans of the CHS beams.

875
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 9. Vertical load-end rotation curves of the CHS beams.

(U1 = U3 = 0; UR1 = UR2 = UR3 = 0). vd = 0.3t f0.2 cr (2)


The tested aluminum alloy specimens in this paper and the previous
research [35] were produced in the same branch; thus, the material where, t is the thickness of the CHS specimens and cr is the elastic local
properties incorporated into the FE models were taken as the mean buckling stress obtained from the eigenvalue buckling analysis.
values of the tensile tests in the literature [35] (with E0 = 68190 MPa,
f0.2 = 285 MPa and n = 28.5). The constitutive equation for 6082-T6 3.2. Element type and mesh size
aluminum alloys could be expressed in the Ramberg-Osgood model
[36,37]: The shell element S4R has been widely adopted for modeling thin-
walled elements—particularly for aluminum alloy specimens. However,
n
according to the ABAQUS manual, the shell elements are not suitable
= + 0.002
E0 f0.2 (1) for CHS specimens with D/t less than 21, and previous research [38] has
indicated that the solid element C3D8I is more suitable than S4R for
According to the measuring results in Section 2.1, the overall geo- modeling CHS specimens. In this study, the accuracy and suitability of
metric imperfections were introduced into the FE models in the shape of C3D8I and S4R in the simulations were evaluated by comparing the
the lowest appropriate elastic buckling mode, which was close to the simulated results for some specimens, as shown in Fig. 13. The simu-
half sinusoid curve. Thus, a linear eigenvalue buckling analysis was lated load–deflection curves of C3D8I were closer to the testing curves
initially performed and the amplitude values v0 was taken as the than the curves of S4R. Although the simulation using C3D8I elements
measured values in Table 1. On the other hand, considering the local took more time than a simulation using S4R elements would have, the
buckling that occurred in the CHS specimens with larger D/t, the local results were more reliable and accurate. Therefore, the selected element
geometric imperfections vd were also incorporated into the FE models for the CHS specimens was taken as C3D8I in the following simulations.
and the values of vd were taken from the Walker expression, which were The mesh size of the aluminum alloy specimens can affect the op-
given in Eq. (2). It should be noted that the length of the CHS beam eration time and ultimate strength of the FE models; thus, a sensitivity
models were taken as their effective length Le. analysis of the mesh size was performed. Eight mesh sizes ranging from
0.5 mm × 0.5 mm to 20 mm × 20 mm were investigated for the

Fig. 10. Vertical load-strain curves of the CHS beams (M-1 and M-3).

876
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 11. Vertical load-strain curves of the CHS beams (Q-1 and Q-4).

aluminum alloy beams CHS-89-5-1800 and CHS-114-4-2200. A com- 4.8%, respectively. The test and simulated results agreed well, in-
parison of the simulated results is presented in Fig. 14. As shown, the dicating that the proposed FE models were sufficiently accurate to
computation time of the FE models increased (geometric progression) predict the ultimate strength, buckling behaviours, and deformation
with a decrease in the mesh size. The differences among the simulated capacity of the CHS beams made of the 6082-T6 aluminum alloy. The
ultimate strengths could be ignored for the models with mesh sizes tested failure modes shown in Fig. 15 were characterised by the residual
ranging from 0.5 mm × 0.5 mm to 5 mm × 5 mm; however, when the deformation of the specimens after unloading; thus, there are small
mesh size exceeded 5 mm × 5 mm, a significant decrease in the ulti- differences in the deformation between the tested and simulated results
mate strength was observed. Thus, considering the computational ef- in Fig. 15.
ficiency and accuracy, the mesh size was taken as 5 mm × 5 mm for the
CHS beams in the following parametric study. Compared with the
3.4. Parametric studies
beams, the deformation of the steel rings could be ignored owing to
their larger Young’s modulus; thus, the sensitivity analysis was not
A series of parametric studies based on the validated FE models
conducted for the steel rings, and the mesh size was uniformly taken as
were performed, with a focus on the variations in the outer diameter
10 mm × 10 mm.
(D) of the cross sections, diameter-to-thickness ratio (D/t), and length of
the specimens. The initial geometric imperfections were integrated into
3.3. Validation of FE models the simulated specimens in the form of the lowest elastic buckling mode
with an amplitude of v0 = Le/1000, according to the measurement
Under the condition that the dimensions of the FE models were results in Section 2.1 and the comparative analysis performed by Zhao
taken as the measured values presented in Table 1, numerical simula- [38]. The mean values of the material properties obtained from the
tions of all 17 tested beams were performed, and the FE ultimate tensile coupon tests [35] (with E0 = 68190 MPa, f0.2 = 285 MPa, and
strength (Nu,FEA) values were compared with the corresponding ex- n = 28.5) were employed in the models of the parametric studies.
perimental values (Nu,test), as shown in Table 2. The ultimate strengths Six different outer diameters of 50 mm, 100 mm, 150 mm, 200 mm,
obtained from the bending tests were slightly higher than the simulated 250 mm and 300 mm were adopted for the simulated specimens. The
values; the difference between them varied from 1% to 3%, with a few diameter-to-thickness ratio (D/t) contributes a lot to the buckling be-
exceptions, e.g. for CHS-114-7-1800A (6.9%) and CHS-89-5-2200B haviors and ultimate strength of the bending specimens; thus, a com-
(4.7%). Figs. 15–17 show typical failure modes, load–deflection curves, bination of 10 different D/t ratios, which varied from 10 to 100 in 10
and load–strain curves; the average difference between the tested spe- intervals, was considered for each outer diameter. The effective length
cimens and the FE models for the deflections and strain was 3.6% and Le of the beam specimens ranged from 500 mm to 5000 mm in 500 mm

Fig. 12. Numerical modeling of the CHS beams subjected to four-point bending.

877
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 13. Comparison of the load-deflections curves using C3D8I and S4R.

Fig. 14. Effect of the mesh size on the simulated results.

Table 2 parametric studies, was expressed in the nondimensional form


Comparison of the ultimate strengths from test and FEA. (Mu,FEA Mel ) and plotted with respect to l in Fig. 18. l is the slen-
Specimens Nu,test (kN) Nu,FEA (kN) Nu,FEA derness of the cross-sections and can be defined as:
Nu,test
l = f0.2 crl (3)
CHS-89-4-1800A 36.62 35.74 0.976
CHS-89-4-1800B 35.84 34.99 0.976 where crl is the critical elastic local buckling stress, which can be ob-
CHS-89-4-1800C 37.64 36.87 0.980 tained using the finite strip software CUFSM. Generally, the cross-sec-
CHS-89-4-2200A 27.84 27.18 0.976
tions with larger l are more likely to fail via local buckling.
CHS-89-5-1800A 49.82 49.07 0.985
CHS-89-5-1800B 47.41 48.16 1.016 For the simulated specimens with the same outer diameters, l
CHS-89-5-2200A 38.64 38.65 1.000 significantly affected the ultimate moment ratios Mu,FEA Mel ; increasing
CHS-89-5-2200B 40.33 38.44 0.953 l reduced the moment ratios. The variation trend of the Mu,FEA Mel
CHS-114-4-1800A 71.10 69.72 0.981 curves was diminished with a decrease in Le, and tended to be more
CHS-114-4-2200A 51.53 50.25 0.975
CHS-114-4-2200B 52.74 50.68 0.961
concentrated as l increased. In contrast, the dispersions of the
CHS-114-5.5-1800A 84.15 83.96 0.997 Mu,FEA Mel curves decreased with an increase in the outer diameters.
CHS-114-5.5-2200A 69.96 69.37 0.992 There was a saltation on the Mu,FEA Mel curves between the specimens
CHS-114-7-1800A 113.83 105.98 0.931 with Le of 500 and 1000 mm, as shown in Fig. 18. The parametric
CHS-114-7-1800B 110.28 108.75 0.986
studies indicated that the specimens with the length of 500 mm were
CHS-114-7-2200A 94.65 91.92 0.971
CHS-114-7-2200B 90.28 89.06 0.987 more likely to fail via local buckling than those with the length of
Mean 0.979 1000 mm or greater, particularly for the specimens with relatively large
COV 0.020 diameters. Thus, the Mu,FEA Mel curves of the 500-mm specimens were
significantly lower than those of the 1000-mm specimens.

intervals. A total of 600 numerical results were generated in the para-


4. Comparison of design provisions for bending strength
metric studies. The numerical database, together with the test results,
were used to assess the accuracy and reliability of the current design
The design provisions of aluminum alloy specimens are available in
specifications, as described in the following sections. For the primary
many specifications, including the commonly used Chinese code GB
parameters, including D, the D/t ratios, and Le, the most common values
50429 (GB) [27], European code Eurocode 9 (EU) [24] and American
in practical engineering were used. To present the simulated results
code AA-2015 (AA) [28]. In recent years, two advanced design meth-
more clearly, the ultimate momentMu,FEA , which was obtained from the
ods—direct strength method (DSM) [29] and continuous strength

878
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 15. Comparison of the failure modes between experimental and numerical results.

method (CSM) [30]—were proposed based on the elastic instability of Otherwise, it is necessary to multiply the origin thickness t with a re-
the gross section and the strain hardening property of the materials, duction coefficient to generate the effective thickness te, which is used
respectively. to calculate the effective section modulus Wex of the cross-section, as
shown in Eq. (4). In Chinese code, the design moment capacity MGB for
the bending specimens should be determined by the resistance of the
4.1. Chinese code GB 50429 (GB)
cross-sections Mu,c and the buckling resistance of the members Mu,b, as
shown in Eq. (5).
The current Chinese code [27] adopts the effective width method
for calculating the effective cross-section of the aluminum alloy speci-
1 2
mens. For the CHS specimens, the whole cross-section will be effective te = t· 0.22· 2
(4)
if the D/t ratios are within the limiting values as listed in Table 3.

879
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 16. Comparison of the load-deflection curves between experimental and numerical results.

MGB = min(Mu,c , Mu,b ) = min( x Wex f0.2 , b Wex f0.2 ) (5) where LT is the reduction factor for lateral torsional buckling; is the
shape factor; Wel is the elastic section modulus of the gross section; Dm is
where 1 and 2 are the coefficients for calculating te ; is the flexibility
the diameter to mid-thickness of CHS specimens.
coefficient; x is plastic adaption coefficient of the cross-section; b is
the stability factor of the bending members.
4.3. American code AA-2015 (AA)

4.2. European code Eurocode 9 (EU) In the last version of American code AA-2010 [39], different design
formulas of the bending specimens are adopted according to their
For the thin-walled specimens, local buckling has a significant effect section types, such as open shapes, closed shapes, round tubes and so
on the buckling behaviors and bearing capacities; thus, four classes of on; however, the current version of American code AA-2015 [28]
cross-sections are defined in Eurocode 9 [24] to identify their de- simplifies these relevant regulations. In AA-2015, the possible buckling
formation capacities under bending. Class 1 and Class 2 sections can modes are divided into the yielding, rupture, local buckling and lateral-
reach their plastic moment resistance and the ultimate bending capacity torsional buckling, and the ultimate strength MAA of the bending spe-
is defined as the yield stress f0.2 multiplied by the plastic modulus Wpl cimens is determined by the least of the available strengths for the
(Wpl = D 2c (1 2 ) 32 for CHS, where c is the distance between the
above buckling modes, as shown in Eq. (10). In view of the complicated
centroid and diameter). Class 3 sections can only reach the elastic calculations for these strengths, only the specific formulas of the
moment resistance due to the local buckling; the elastic plastic modulus bending strengths for yielding and rupture are presented in this paper,
Wel is used to calculate the bending capacity. Class 4 sections are those as shown in Eqs. (11) and (12), and the other strengths are available in
in which the attainment of the elastic moment is prevented by the the Section F.3 to F.5 of AA-2015.
premature local buckling; the elastic modulus of the effective sections
Weff is adopted. The ultimate bending capacities of the CHS specimens MAA = min(Mnp , Mnu, Mnlb, Mnmb) (10)
are given by Eqs. (6)–(8). The cross-section slenderness parameter is
Mnp = min(ZFcy , 1.5St Fty , 1.5Sc Fcy ) (11)
defined as Eq. (9) and the limiting values for 1, 2 and 3 are listed in
Table 4. Mnu = ZFtu k t (12)
MEU = LT Wel f0.2 (6) where Mnp , Mnu , Mnlb and Mnmb are the nominal bending strength for the
limit state of yielding, rupture, local buckling and lateral-torsional
Wpl Wel 1 (for Class 1) buckling, respectively; Z is the plastic modulus; St and Sc are the section
Wpl Wel 1 < 2 (for Class 2) modulus on the tension and compression side of the neutral axis, re-
=
3,u 2 < 3 (for Class 3) spectively; Fcy is the compressive yield strength; Fty is the tensile yield
Weff Wel < (for Class 4) strength; Ftu is the tensile ultimate strength; k t is the tension coefficient.
3 (7)
4.4. Direct strength method (DSM)
3 Wpl
3,u = 1, or 1 + 1
Wel (8)
3 2 The direct strength method (DSM) [29] was proposed by Schafer
and Pekoz in 1998, aiming at the cold-formed carbon steel members
= 3 Dm t (9) failed by local or distortional buckling. The biggest advantage of DSM is

Fig. 17. Comparison of the load-strain curves between experimental and numerical results.

880
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 18. Numerical results of ultimate moment versus effective length.

Table 3 Table 4
Limiting values of D/t ratios in GB 50429 [27]. Limiting values of cross-section slenderness parameters [24].
Hardening degree Non-welded Welded Material type Without welds With welds

Weak hardening 50·(240 f0.2 ) 35·(240 f0.2 ) 1 2 3 1 2 3


Strong hardening 35·(240 f0.2 ) 25·(240 f0.2 )
Class A 11 16 22 9 13 18
Class B 13 16.5 18 10 13.5 15

= 250 f0.2 , f0.2 in N/mm2.


that the computational procedure of the ultimate bearing capacity can
be much concise for the members with complex cross-sections com-
pared to the effective section method (ESM), which is adopted in many
design specifications. Zhu and Young [16,40] have assessed the ap- the members with closed sections. According to DSM, the ultimate
plicability and accuracy of DSM on aluminum alloy specimens sub- bending strength of aluminum alloy beams is governed by the minimum
jected to compression or bending and adjusted the key coefficients for of the local (Mnl ), distortional (Mnd ) and global (Mne ) buckling strength.

881
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

In this study, the specimens with closed section CHS were investigated csm 4.44 × 10 - 3 csm 0.5 u
and distortional buckling did not occur during the bending tests; = 4.5
for c 0.3 but min 15,
y c y y (17)
therefore, Mnd was not considered in this section and the ultimate
bending capacity MDSM was calculated by Eqs. (13)–(15). 0.224 1
csm
= 1 for 0.3 < c 0.6
MDSM = min(Mne, Mnl) (13) y c
0.342
c
0.342
(18)

Mne for d 0.713 fu f0.2


Esh =
Mnl = 0.5 (19)
1 0.15 ( ) ( )
Mcrl 0.3
Mne
Mcrl 0.3
Mne
Mne for d > 0.713
(14)
u 0.2

u = 0.13(1 f0.2 fu ) + 0.06 (20)


Mcre for Mcre < 0.56My
2
Esh Wel Wel
Mne =
10
9
My (1
10My
36Mcre ) for 0.56My Mcre 2.78My Mcsm = Mpl 1 +
E Wpl
csm
y
1 1
Wpl
csm
y
for c 0.3
My for Mcre > 2.78My (15) (21)
where My = f0.2 Sg is the yield moment of the cross-section; Sg is the csm
Mcsm = Mel for 0.3 < 0.68
cross-sectional modulus referenced to the extreme fiber in first yield; y
c
(22)
d = Mne Mcrl is the slenderness of the cross-sections in DSM; Mcrl is
the critical elastic local buckling moment; Mcre is the critical elastic where = 0.33 is the Poisson’s ration; csm is the CSM limiting strain;
lateral-torsional buckling moment. 0.2 = f0.2 E0 is the yield strain; u is the strain at ultimate tensile stress;
Esh is the strain hardening modulus; fu is the ultimate tensile strength;
4.5. Continuous strength method (CSM) Mpl = f0.2 Wpl and Mel = f0.2 Wel are the plastic and elastic moment ca-
pacity of the cross-section, respectively; Wpl and Wel are the plastic and
Aluminum alloys exhibit a high degree of material nonlinearity; elastic section modulus, respectively.
their post-yield properties, such as strain hardening and ductility, have
a significant effect on the buckling behaviors and ultimate strengths. In 4.6. Comparisons and analysis
view of this, the group of Gardner [30] proposed an advanced design
approach—continuous strength method (CSM), which takes into con- The ultimate bending capacities (Mu,test FEA ) obtained from the tests
sideration the beneficial influence of the material strain hardening by a and simulations were compared with the design strengths predicted
continuous stress–strain curves. CSM is a deformation (strain) based using three current specifications and two advanced design methods
design method including two key components: (1) a base curve that (Mu,GB , Mu,EU , Mu,AA , Mu,DSM , and Mu,CSM ). The mean values and coeffi-
defines the continuous relationship between the cross-section de- cients of variation (COVs) of the moment ratios are presented in
formation capacity and local slenderness; (2) a bi-linear (elastic, linear Table 3. Comparisons between Mu,test FEA and the design strengths
hardening) material model that allows for the influence of strain predicted by the five design provisions are shown in Figs. 19–23. For
hardening. The cross-section slenderness is defined as c = f0.2 cr , the design-strength calculations, the mean values of the material
where f0..2 is the nominal yield stress and cr is critical elastic buckling properties were adopted rather than the standard values, and the partial
capacity of the cross-section (Eq. (16)). For the aluminum alloy beams, factors of the design provisions were set as 1.0 uniformly.
the base curve defining the deformation capacity of CHS is given in Eqs. The moment ratios of the test and numerical results to the predic-
(17) and (18) [41] and the related parameters of the bi-linear material tions are presented in Table 3, together with the COVs. Overall, on
model are determined by Eqs. (19) and (20). Having determined the average, all the five design provisions were conservative in calculating
key parameters of CSM, the ultimate bending capacity Mu,CSM can the ultimate bending capacity of the aluminum alloy CHS beams. The
subsequently calculated by the Eqs. (21) and (22), which is available in Chinese code GB 50429 generally underestimated the beam strengths
[41]. by approximately 40%, with the largest scatter among the design pro-
visions (Mu,test FEA Mu,GB = 1.395 and COV = 0.101). The European
E0 2t
cr = code Eurocode 9 and American code AA-2015 more accurately pre-
3(1 2) D (16) dicted the ultimate strength, with mean moment ratios of 1.297 and

Fig. 19. Comparison of experimental and numerical results with design capacities by GB 50429.

882
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 20. Comparison of experimental and numerical results with design capacities by Eurocode 9.

1.257 and corresponding COVs of 0.091 and 0.071, respectively. A variation ranges of the moment ratios in the Chinese code was the
significant improvement in the accuracy of the DSM is observed in largest among the design provisions, owing to the reduction of the
Table 3; the mean moment ratio and COV were 1.155 and 0.056, re- cross-section thickness. For the specimens with large D/t ratios or l
spectively. The CSM provision provided the best prediction of the values, their thickness should be reduced by multiplying by a reduction
bending capacity (Mu,test/FEA /Mu,CSM = 1.117 and COV = 0.048), with an coefficient, yielding a smaller Mu,GB and larger moment ratios. For the
accuracy improvement of approximately 28% compared with GB CSM, the moment-ratio points were more concentrated on the plot
50429. compared with the other provisions, and the variations were relatively
In the Chinese code, only the elastic section modulus Wel is em- stable, corresponding to the smallest COV, as shown in Table 5.
ployed to calculate the bending capacity, and the section plastic In this study, the specimens with small D/t (≤20) and l (≤0.3)
adaption coefficient x is taken as 1.0. Hence, the development of values were assigned to the Class 1 and Class 2 sections of Eurocode 9;
plastic deformation is not considered for the cross sections. Thus, the the cross sections can reach their plastic-deformation limits, and the
moment ratios of the experimental or simulated results to the predicted bending moments meet the ultimate values. The bending strengths
values were the largest for GB 50429. For the other four design provi- calculated using the plastic section modulus Wpl were largely consistent
sions, the cross sections were divided into several classes to identify the with the testing status; thus, in Fig. 20, the moment ratios of the spe-
extent to which the plastic deformation could develop—particularly for cimens with a smaller l are closer to the red line (predicted results)
the CSM, which also considers the strain hardening of the sections after than those with a larger l (> 0.3).
yielding. Therefore, the bending capacities predicted using these spe-
cifications were more accurate than those predicted using the Chinese
code. 5. Reliability assessment of design provisions
Figs. 19–23 shows that all the simulated bending capacities were
larger than the predicted results. Except for the CSM, the moment-ratio In this section, a reliability analysis was carried out to assess the
points of the design provisions exhibited a trend of first decreasing and safety levels of the current design specifications (GB 50429, Eurocode 9
then increasing with an increase in l , and the turning points were lo- and AA-2015) and the recently proposed design methods (DSM and
cated between l = 0.35 and l = 0.50 . In contrast, for the CSM, the CSM) for the aluminum alloy beams. Meanwhile, an advanced method
moment ratios declined with an increase in l , and the declining rate for determining the distributions of model errors (ME) was introduced
tended to decrease with an increase in the cross-section diameter. The in this paper.

Fig. 21. Comparison of experimental and numerical results with design capacities by AA-2015.

883
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 22. Comparison of experimental and numerical results with design capacities by DSM.

5.1. Resistance and load variables Table 5


Comparisons of the test and FEA results with predicted strengths.
Three important variables of the resistance—the material strength, Statistic Mu,test FEA Mu,test FEA Mu,test FEA Mu,test FEA Mu,test FEA
geometric properties, and ME—were investigated in this study. The parameters MGB MEU MAA MDSM MCSM

statistical parameters of the material strength for 6082-T6 aluminum


alloys were obtained from the results of tensile tests reported in the Mean 1.395 1.297 1.257 1.155 1.117
COV 0.101 0.091 0.071 0.056 0.048
literature [35]. The measurement results for the section dimensions, Distribution Gumbel Weibull lognormal Gumbel lognormal
which are presented in Table 1, were utilised to calculate the geometric
properties. Table 6 presents the mean values and COVs of the foregoing
two variables. Their probability distributions were validated via stan- Table 6
dard Kolmogorov–Smirnov goodness-of-fit tests at a 5% significance Statistical parameters of material and geometries.
level.
Variables Mean COV Distribution
The ME is defined as the ratio of the experimental and simulated
results to the predicted ultimate strengths. The mean values and COVs D0/D, t0/t 0.993 0.014 Normal
of the ME for the five design provisions are presented in Table 5. The f0.2 (MPa) 301.89 0.095 Normal
distribution types of the ME can significantly affect the reliability-
analysis results [42]; thus, five common distributions [43]—normal,
lognormal, Gumbel, Weibull, and gamma—were investigated to iden- by the distance between the 1:1 line and the “lower tails” of the CDF−1
tify the optimal ones for the various provisions considered in this study. curves; thus, the lognormal distribution was adopted for AA-2015. The
Fig. 24 shows the probability density histograms and the inverse cu- subsequent analysis revealed that the Gumbel distribution was the
mulative distribution functions (CDF−1) for AA-2015. Five fitted dis- optimal distribution for GB 50429 and DSM, whereas the Weibull and
tribution curves are plotted against the density histograms in Fig. 24(a); lognormal distributions were optimal for Eurocode 9 and CSM, re-
the Kolmogorov–Smirnov test results indicated that no distributions spectively. The distribution types of the ME for the design provisions,
were rejected at the significance level of 5%. Fig. 24(b) shows a com- together with the mean values and COVs (as shown in Table 5), were
parison between the CDF−1 of the five distributions and the 1:1 line. In directly applied in the following reliability analysis.
this study, the optimal fitted distributions of the ME were determined The dead load (D k ) and live load (Lk ) were used to calculate the

Fig. 23. Comparison of experimental and numerical results with design capacities by CSM.

884
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Fig. 24. Probability distribution fitting for AA-2015.

design load effect, together with the resistance factor ( ). For the alu- analysis.
minum alloy specimens, the considered design provisions adopt dif- The advanced first-order second-moment method was used to cal-
ferent load combinations and values [43]: (1) GB 50429 uses culate the reliability indices of the design provisions by using MATLAB.
= 0.833 and a load combination of 1.3D k + 0.98Lk ; (2) Eurocode 9 The analysis results are presented in Table 7. Among the five design
uses = 0.91 and a load combination of 1.35D k + 1.5Lk ; (3) AA-2015 provisions, Eurocode 9 had the largest reliability index, followed by GB
uses = 0.90 and a load combination of 1.2D k + 1.6Lk . The load 50429. The DSM and CSM had the third- and fourth-lowest safety le-
combinations and values of the DSM and CSM are not available in the vels, respectively, and the reliability level of AA-2015 was the lowest.
literature or in design specifications; thus, the specified and load However, the average reliability indices ( ) of GB 50429, Eurocode 9,
combination of GB 50429 were adopted for the DSM and CSM, con- and AA-2015 were higher than the corresponding target values
sidering that the tested aluminum alloy specimens were produced in [ ]GB = 3.7 , [ ]EU = 3.8, and [ ]AA = 2.5 , indicating that the design
China. The distributions and statistical parameters of D k and Lk were specifications of these three codes are reliable and safe for the alu-
set according to recommendations in the literature [43]. minum alloy beams. The variations of the reliability indices ( ) are
shown in Fig. 25, with respect to . For all five design provisions, first
5.2. Reliability-analysis results increased and then decreased, and the turning points of the load ratio
were located at = 0.5. The variation ranges of for the DSM and CSM
Three classes of safety levels are defined in the Chinese code GB were significantly larger than those for the three design codes, in-
50429: the major class, common class, and minor class. For the alu- dicating that the reliability levels of the DSM and CSM were more
minum alloy structures, the target reliability indices [ ] corresponding susceptible to the load ratios than those of the design codes.
to these safety classes were set as 4.2, 3.7, and 3.2 [27], respectively.
Similar rules are specified in the European code Eurocode 9; the [ ] 6. Conclusions
values were set as 4.3, 3.8, and 3.3 for the three safety classes (CC3,
CC2, and CC1, respectively) [44]. In this study, the second safety class We experimentally investigated 17 CHS aluminum alloy specimens
was investigated for the aluminum alloy structures, and thus the target subjected to four-point bending. The ultimate bending strength and
reliability indices of GB 50429 and Eurocode 9 were set as [ ]GB = 3.7 buckling modes were examined, and load–deflection curves and
and [ ]EU = 3.8, respectively. For AA-2015, the regulations regarding load–strain curves were acquired. FE models were developed using
[ ] are not as strict as those for GB 50429 and Eurocode 9, and the ABAQUS and validated via comparisons between the numerical pre-
values of [ ] were uniformly set as 2.5 [28]; thus, the target reliability dictions and test results. An extensive parametric study was performed
index of AA-2015 was set as [ ]AA = 2.5 . The [ ] values of the DSM and to evaluate the effects of key parameters, including the section di-
CSM were not specified in this study, and the safety levels were eval- mensions, diameter-to-thickness ratio, and specimen length, on the ul-
uated by comparing the reliability-analysis results with those of the timate strength and buckling characteristics. The experimental and
three design codes. Another key parameter was the load ratio ( ), numerical results were employed to evaluate the applicability of the
which was defined as the live load divided by the dead load and was set bending-strength provisions from the Chinese code GB 50429,
as 0, 0.25, 0.5, 1.0, 1.5, 2.0, 2.5, 3.0, 3.5, and 4.0 in the reliability European code Eurocode 9, American code AA-2015, DSM, and CSM.

Table 7
Reliability indices of the aluminum alloy CHS beams.

Design provisions Load ratio = L D [ ]

0.01 0.25 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

GB50429 4.40 4.49 4.53 4.53 4.48 4.41 4.35 4.29 4.25 4.21 4.39 3.7
Eurocode 9 4.23 4.57 4.73 4.73 4.63 4.55 4.49 4.44 4.41 4.38 4.51 3.8
AA-2015 3.82 4.17 4.23 4.06 3.92 3.83 3.77 3.73 3.69 3.67 3.89 2.5
DSM 4.73 4.85 4.87 4.66 4.42 4.25 4.14 4.05 3.98 3.93 4.33 —
CSM 4.73 4.87 4.87 4.60 4.35 4.17 4.05 3.96 3.90 3.84 4.28 —

885
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influ-
ence the work reported in this paper.

Acknowledgements

This research program is financially supported by National Natural


Science Foundation of China (Grant no. 51978208). Special thanks to
the Key Lab of Structures Dynamic Behavior and Control at Harbin
Institute of Technology for providing the experimental sites and mea-
suring instruments.

References

[1] Mazzolani FM. Aluminum alloy structures. Second edition London: Spon Press;
Fig. 25. Reliability indices of the five design provisions as the function of load 1994.
ratio α. [2] Mazzolani FM. Competing issues for aluminium alloys in structural engineering.
Progr Struct Eng Mater 2004;6(2):185–96.
[3] Soetens F. Aluminium structures in building and civil engineering applications.
The safety levels of the five design provisions were assessed via a re- Struct Eng Int 2010;4:430–5.
[4] De Matteis G, Mazzolani G, Brando FM. Pure aluminium: an innovative material for
liability analysis. The detailed findings of this study are summarised as structural applications in seismic engineering. Constr Build Mater
follows: 2012;26(1):677–86.
[5] Sharp ML. Behaviour and design of aluminium structures. New York: McGraw-Hill;
1993.
(1) CHS aluminum alloy specimens subjected to four-point bending [6] C. Dumont, H.N. Hill, The lateral instability of deep rectangular beams, NACA Tech.
with larger diameter-to-thickness ratios (D t 20 ) were more likely Note 601. 1937.
to fail via local buckling in the post-loading stage than thicker [7] C. Dumont, H.N. Hill, Lateral stability of equal flanged aluminum alloy I-beams
subjected to pure bending, NACA Tech. Note 770. 1940.
specimens (D t < 20 ), which failed via overall buckling. When the
[8] Hill HN. The lateral instability of unsymmetrical I-beams. J Aero Sci 1942;9:175.
local buckling occurred, the ultimate load suddenly decreased, and [9] Clark JW, Jombock JR. Lateral buckling of I-beams subjected to unequal end mo-
the deformation capacity was lower than that for the specimens that ment. J Eng Mech Div, ASCE 1957;83(3):1–19.
[10] F.M. Mazzolani V. Piluso Evaluation of the rotation capacity of steel beams and
failed via overall buckling. The mid-span deflection and end rota-
beam-columns, Proc., 1st COST C1 Workshop 1992 Strasbourg 517 529.
tion capacity of the CHS aluminum alloy beams were inversely [11] Mazzolani FM, Piluso V. Prediction of the rotation capacity of aluminum alloy
proportional to the ultimate moment capacity, and their growth beams. Thin-Walled Struct 1997;27(1):103–16.
rates were proportional to the ultimate strength. For the same cross [12] Mazzolani FM, Cappelli M, Spasiano G. Plastic analysis of aluminium alloy members
in bending. Aluminum 1985:61.
sections, longer specimens corresponded to a higher ultimate strain. [13] Zhu JH, Young B. Design of aluminum alloy flexural members using direct strength
(2) According to the validated FE models considering the material and method. J Struct Eng ASCE 2009;135(5):558–66.
geometric nonlinearity, the simulated results of parametric study [14] Su M, Young B, Gardner L. Deformation-based design of aluminium alloy beams.
Eng Struct 2014;80:339–49.
indicated that the decrease in the D t ratios increased the ultimate [15] Su M, Young B, Gardner L. Continuous beams of aluminum alloy tubular cross-
moment ratios Mu,FEA Mel for the specimens with the same dia- section – part I: tests and model validation. J Struct Eng ASCE
meter; the moment ratios Mu,FEA Mel increased with the increment 2015;141(9):04014232.
[16] Su M, Young B, Gardner L. Continuous beams of aluminum alloy tubular cross-
of the length for the specimens with the same cross section. section – part I: parametric study and design. J Struct Eng ASCE
Mu,FEA Mel hardly changed with the increase of the D t ratio; there 2015;141(9):04014233.
was an obvious saltation for Mu,FEA Mel between the specimens with [17] Moen LA, Hopperstad OS, Langseth M. Rotational capacity of aluminum beams
under moment gradient. I: experiments. J Struct Eng 1999;125(8):910–20.
lengths of 500 and 1000 mm, because local buckling was more
[18] Moen LA, Matteis GD, Hopperstad OS. Rotational capacity of aluminum beams
likely to occur in the shorter specimens. under moment gradient. II: numerical simulations. J Struct Eng 1999;125(8):921–9.
(3) The accuracies of the design provisions were evaluated. All five [19] Moen LA, Langseth M, Hopperstad OS. Elastoplastic buckling of anisotropic alu-
minium plate elements. J Struct Eng 1998;124(6):712–9.
design provisions were conservative in predicting the ultimate
[20] Wu YG, Zhang QL. Numerical and experimental study on flexural-torsional buckling
bending capacity of the aluminum alloy beams. The CSM provided coefficient of aluminum beams. J Build Struct 2006;27(5):1–8. (in Chinese).
the most accurate mean values, with the smallest COV of the ulti- [21] Guo XN, Shen ZY, Li YQ, et al. Theoretical and experimental research on aluminum
mate strength (Mu,test/FEA Mu,CSM = 1.117and COV = 0.048). GB alloy beams. J Build Struct 2007;28(6):129–35. (in Chinese).
[22] X.N. Guo Theoretical and experimental research on the aluminum alloy structure
50429 provided the most conservative results, with the largest COV members 2006 Thesis, Tongji University Doctor (In Chinese).
(Mu,test/FEA /Mu,GB = 1.395and COV = 0.101). The predicted results [23] Wang YQ, Yuan HX, Shi YJ. et at., Lateral-torsional buckling resistance of alumi-
for the CSM decreased with an increase in the cross-section slen- nium I-beams. Thin-Walled Struct 2012;50:24–36.
[24] BS EN 1999-1-1, Eurocode 9: Design of Aluminum Structures—Part 1-1: General
derness l , whereas the results for the other four design provisions structural rules, European Committee for Standardization (CEN), Brussels, 2007.
first increased and then decreased. [25] Chen Y, Feng R, Xu J. Flexural behaviour of CFRP strengthened concrete-filled
(4) The reliability analysis indicated that Eurocode 9 provided the most aluminium alloy CHS tubes. Constr Build Mater 2017;142:295–319.
[26] Chen Y, Feng R, Gong WZ. Flexural behavior of concrete-filled aluminum alloy
reliable results (reliability index of EU = 4.51), followed by GB circular hollow section tubes. Constr Build Mater 2018;165:173–86.
50429 ( GB = 4.39). The reliability index of AA-2015 was the lowest [27] MOHURD, Code for design of aluminum structures GB 50429-2007, China Planning
among the five design provisions ( AA = 3.89 ). All the values sa- Press, Beijing, 2007 (in Chinese).
[28] Aluminum Association, Aluminum Design Manual, The Aluminum Association,
tisfied the corresponding target reliability indices, i.e. [ ]EU = 3.8, Washington, DC, 2015.
[ ]GB = 3.7 , and [ ]AA = 2.5 for Eurocode 9, GB 50429, and AA- [29] B.W. Schafer, T. Pekoz, Direct strength prediction of cold-formed steel members
2015, respectively. DSM and CSM had the third- and fourth-lowest using numerical elastic buckling solutions. In: Proceedings of the 14th international
specialty conference on cold-formed steel structures. Rolla, Mo: University of
safety levels, and were more sensitive to the load ratios than the
Missouri-Rolla; 1998. p. 69–76.
other provisions. [30] Gardner L. The continuous strength method. Proc ICE – Struct Build
2008;161(3):127–33.

886
Y. Zhao and X. Zhai Structures 26 (2020) 870–887

[31] Zhao YZ, Zhai XM, Sun LJ. Test and design method for the buckling behaviors of [38] Zhao YZ, Zhai XM, Wang JH. Buckling behaviors and ultimate strength of 6082–T6
6082–T6 aluminum alloy columns with box-type and L-type sections under ec- aluminum alloy columns with square and circular hollow sections under eccentric
centric compression. Thin-Walled Struct 2016;100:62–80. compression – Part II: parametric study, design provisions and reliability analysis.
[32] F.M. Mazzolani, Residual Stress Tests Alu-Alloy Austrian Profiles, ECCS Committee, Thin-Walled Struct 2019;143:106208.
Brussels, 1975, Technical Report, Doc 16-75-1. [39] Aluminum Association, Aluminum Design Manual, The Aluminum Association,
[33] Kiymaz G, Coskun E, Cosgun C. Behavior and design of seam-welded stainless steel Washington, DC, 2010.
circular hollow section flexural members. J Struct Eng 2007;133(12):1792–800. [40] Zhu JH, Young B. Aluminum alloy tubular columns–Part II: Parametric study and
[34] Simulia, ABAQUS Standard, User’s Manual, Version 6.10, Rhode Island, USA, 2010. design using direct strength method. Thin-Walled Struct 2006;44:969–85.
[35] Zhao YZ, Zhai XM, Wang JH. Buckling behaviors and ultimate strengths of 6082–T6 [41] Buchanan C, Gardner L, Liew A. The continuous strength method for the design of
aluminum alloy columns under eccentric compression – Part I: Experiments and circular hollow sections. J Constr Steel Res 2016;118:207–16.
finite element modeling. Thin-Walled Struct 2019;143:106207. [42] X.M. Zhai, M.G. Stewart, Structural reliability analysis of reinforced grouted con-
[36] SteinHardt O. Aluminium constructions in civil engineering. Aluminium crete block masonry walls in compression, Eng. Struct. 32 (2010) 106-114.
1971;47:131–9. [43] Zhao YZ, Zhai XM. Reliability assessment of aluminum alloy columns subjected to
[37] Ramberg W, Osgood WR. Description of stress strain curves by three Parameters: axial and eccentric loadings. Struct Saf 2018;70:1–13.
technical note 902. Washington, DC: National Advisory Committee for Aeronautics; [44] BS EN 1990-2002, Eurocode: Basis of Structural Design, European Committee for
1943. p. 1–12. Standardization (CEN), Brussels, 2002.

887

You might also like