Sie sind auf Seite 1von 10

Journal of Constructional Steel Research 64 (2008) 505514

www.elsevier.com/locate/jcsr

Experimental study of beamcolumn behaviour of steel single angles


Yi Liu , Linbo Hui
Department of Civil and Resource Engineering, Dalhousie University, Halifax, NS, Canada
Received 24 July 2007; accepted 9 November 2007

Abstract
Twenty-eight rolled steel single angle specimens were tested to investigate their response when required to carry axial compressive loading at
various end eccentricities. Results suggested that when eccentrically loaded with respect to the major principal axis, there is a critical eccentricity
below which any consequent reduction in the ultimate load is marginal. In contrast, as eccentricity of loading with respect to the minor principal
axis is increased, reduction of the ultimate load is more pronounced and no similar critical eccentricity can be identified. Test results, when
compared with the corresponding values as determined from the design equations suggested by Adluri and Madugula (1992), in AISC Specification
2000 and AISC Specification 2005, indicated that the former two methods give a conservative estimate of the ultimate compressive capacity of
single angles. This conservatism is more pronounced for specimens subjected to eccentric loading with respect to the major principal axis than
that resulting from eccentric loading with respect to the minor principal axis. Although intended for doubly symmetric sections, the third method
provides improved capacity estimates of single angles.
c 2007 Elsevier Ltd. All rights reserved.

Keywords: Single angle; Eccentric compression; Beamcolumn; Major-axis bending; Minor-axis bending; AISC specification

1. Introduction
Single steel angles are extensively used in a variety of
structures such as steel joists and trusses, latticed transmission
towers and antenna-supporting towers. In practical applications,
single angles are usually either welded or bolted to other
structural members by one leg. This connection detail leads
to eccentric loading and bi-axial bending about principal
or geometric axes of the member. Due to the complexity
associated with combined axial stress and bending coupled
with typical asymmetric characteristics of angles, previous
research into the response of these members has mainly
been of an experimental nature supplemented by numerical
modeling.
The design of single angles has evolved into two major
approaches in North America. Developed by the lattice tower
industry, one approach treats eccentrically loaded single angles
as simple columns. To account for both end eccentricity
and end restraint, empirical equations are used wherein a
modified slenderness ratio is incorporated in an appropriate
Corresponding author. Tel.: +1 902 494 1509; fax: +1 902 494 3108.

E-mail address: yi.liu@dal.ca (Y. Liu).


c 2007 Elsevier Ltd. All rights reserved.
0143-974X/$ - see front matter
doi:10.1016/j.jcsr.2007.11.002

column formula [1,2]. In these provisions, the influence of load


eccentricity is controlled by specifying a range of bolt hole
locations resulting in what is referred to as normal framing
eccentricity.
The second approach evaluates the capacity of eccentrically
loaded single angles using equations that include the effects
of the interaction between axial load and bending. Used in
general building construction involving angles, this approach
is recommended in AISC Specification 2000 [3], AISC
Specification 2005 [4] and the current Canadian steel
design standard CSA S16-01 [5]. In contrast to the simple
column design approach, axial force and moment interaction
equations are used to reflect the more realistic load response
characteristics of angles. As suggested in AISC Specification
2000 the interaction of flexure and axial compression applicable
to specific locations on a cross-section shall be limited
according to:


Mr y
Pr
Pr
8
Mr x
1.0
For
0.2,
+
+
c Pn
c Pn
9 b Mnx
b Mny
(1a)


Mr y
Pr
Pr
Mr x
For
< 0.2,
+
+
1.0 (1b)
c Pn
2c Pn
b Mnx
b Mny

506

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

where Pr and Pn are the required compressive strength and


the nominal compressive strength ; Mr is the required flexural
strength and Mn is the nominal flexural strength; c and b are
the resistance factors for compression and flexure, and; x and y
refer to the major and minor principal axes. AISC Specification
2005 for structural steel buildings adopts the above equations
for the design of singly symmetric members in flexure and
compression.
Adluri and Madugula [6] pointed out that these interaction
equations are derived mainly from the research on doubly
symmetric W -sections and the values for moment interaction
factors are obtained from the experimental results of
82 sidesway frames composed of doubly symmetric W sections [7]. Their applicability to eccentrically loaded
single angle members which are either mono-symmetric or
asymmetric is questionable. Based on a comparison study of
71 test results from three different sources [810], they showed
that the equations in AISC Specification 2000 are highly
conservative and proposed modified equations as follows:


Mr y
Pr
2
Mr x
Pr
1.0
For
0.5,
+
+
c Pn
c Pn
3 b Mnx
b Mny
(2a)


Mr y
Mr x
Pr
Pr
< 0.5,
+
+
1.0 (2b)
For
c Pn
2c Pn
b Mnx
b Mny
where all terms are defined as before.
The effectiveness of the equations in AISC Specification
2000 and those suggested by Adluri and Madugula [6] was
examined by Sakla [11] using an experimental database
containing 133 test results of both equal and unequal-leg
angles. Sakla concluded that AISC Specification 2000 provides
a better prediction of the load carrying capacity of equal-leg
single angle members than it does for unequal-leg single angle
members. It underestimates all unequal-leg test results and
overestimates capacities of some equal-leg angles. Sakla also
pointed out that, while the equations proposed by Adluri and
Madugula [6] provide an improved average test-to-prediction
ratio of the load carrying capacity for equal and unequalleg angles, the standard deviation of the ratio remains almost
unchanged. The modified equations ((2a) and (2b)) fell short
of achieving the target reliability index when used to predict
the capacity of equal-leg single angles. Earls [12] summarized
the research findings reported by Madugula et al. [13] and Earls
and Galambos [14]. He suggested that AISC Specification 2000
for single angle flexural design is, in most cases, unnecessarily
conservative. He also identified the situations where the AISC
Specification 2000 overestimates predicted capacities for single
angle beams.
In Canada, while the interaction approach is recognized
in the current Canadian steel design standard CSA S16-01,
the axial force and moment interaction factors are different
from the AISC Specifications [3,4]. Additionally, CSA S16-01
does not include explicit provisions for calculating the nominal
flexural strength for mono-symmetric or asymmetric angles, but
rather, it directs the readers to the SSRC Guide [15] for design
methods.

Fig. 1. Angle cross-section and definitions.

Although some previously mentioned experimental research


on eccentrically loaded angles was carried out on angles with
various slenderness ratios, the exact end eccentricities and
boundary conditions used in these experiments were not clearly
reported making an accurate interpretation of the test results
difficult. In addition, where end eccentricities were specified,
most eccentricities were within the so-called normal framing
eccentricity range and the effect of large eccentricities was not
clear.
In the research described herein, twenty-eight single
angles were tested under various compressive loadings and
eccentricity conditions [16]. Being the most commonly used
in practice, only equal-leg single angles were considered in this
study. Test results serve to augment the existing test database
and to further examine the effectiveness of the interaction
equations as are presently applicable in design.
2. Test specimens
Twenty-eight L51 51 6.4 single steel angles were
tested to ultimate under compressive loading applied at various
eccentricities. The salient dimensional, material, loading, and
test response details of all specimens are summarized in
Table 1. Specimens were categorized into three series based
on nominal lengths of 900, 1200, and 1500 mm. The effective
length for each specimen as listed in the table includes a
consideration of the thickness of bearing plates onto which
the specimen was tack welded. Eccentricities, ex and e y , with
respect to the major and minor principal axes, x and y, are
as illustrated in Fig. 1. The initial out-of-straightness along
the length was measured for each specimen with respect to
the principal axes using a technique similar to that reported
by Adluri and Madugula [17]. All measurements were taken
from a datum formed by a piano wire tightly stretched over
a length parallel to an angle. The maximum initial out-ofstraightness for all specimens was less than L/1500 with an
average value of L/2100 and a coefficient of variation (C.O.V)
of 8.7%. Nominally identical tests were repeated to ensure that
the testing procedure delivered the consistent results.

507

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514


Table 1
Details of test specimens and results
No.

Specimen ID#

Length K L (mm)

K L/r y

Fy (MPa)

E (GPa)

ex (mm)

e y (mm)

P (kN)

Failure mode

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28

E900-1
E900-2
E900-3
E900-4
E900-5
E900-6
E900-7
E900-8
E900-9
E900-10
E1200-1a
E1200-1b
E1200-2
E1200-3a
E1200-3b
E1200-4
E1200-5
E1500-1a
E1500-1b
E1500-1c
E1500-2
E1500-3
E1500-4
E1500-5
E1500-6
E1500-7
E1500-7
E1500-7

941
945
948
940
940
940
939
939
944
944
1243
1247
1243
1243
1243
1244
1246
1548
1546
1548
1546
1546
1546
1544
1544
1552
1550
1544

94.8
95.2
95.5
94.7
94.7
94.7
94.6
94.6
95.1
95.1
125.2
125.6
125.2
125.2
125.2
125.3
125.3
155.9
155.7
155.9
155.7
155.7
155.7
155.5
155.5
156.3
156.1
155.5

330.0
330.0
330.0
330.0
330.0
330.0
347.9
347.9
347.9
347.9
330.0
330.0
330.0
347.9
330.0
330.0
347.9
347.9
347.9
347.9
347.9
347.9
347.9
347.9
347.9
347.9
347.9
347.9

199.9
199.9
199.9
199.9
199.9
199.9
202.5
202.5
202.5
202.5
199.9
199.9
199.9
202.5
199.9
199.9
202.5
202.5
202.5
202.5
202.5
202.5
202.5
202.5
202.5
202.5
202.5
202.5

0.0
4.2
10.0
16.8
29.4
50.4
0.0
0.0
0.0
0.0
4.2
4.2
10.0
16.8
16.8
29.4
50.4
0.0
0.0
0.0
8.4
16.8
29.4
50.4
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
8.4
16.8
29.4
50.4
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
8.4
16.8
29.4
50.4

117.0
110.9
105.3
84.5
59.0
46.1
65.6
46.4
34.4
23.3
68.1
69.7
66.7
66.4
71.3
57.0
42.5
45.3
41.1
42.8
42.9
41.3
40.1
37.0
35.2
27.8
23.4
18.0

FB
TFB
TFB
TFB
TFB
TFB
FB
FB
FB
FB
FB
FB
TFB
TFB
TFB
TFB
TFB
FB
FB
FB
FB
TFB
TFB
TFB
FB
FB
FB
FB

Measured dimensions of specimens varied from the nominal


dimensions as listed in CSA S16.01 by less than 2.0% for the
leg width and by less than 3.0% for the leg thickness. The
average measured-to-nominal ratios for leg width and thickness
dimensions were close to 1.00 and had C.O.Vs less than 1.3%.
Test specimens were cut from CSA G40.21 grade 300W [5]
angles. Nine tension coupons taken from each of two
stock shipments were fabricated and tested to determine
the mechanical properties of the angle steel according to
ASTM [18]. The average value of yield stress, Fy , and Youngs
modulus, E, as obtained from coupon tests are listed in Table 1
for each specimen.
3. Test set-up
The test set-up is shown in Fig. 2. An Instron hydraulic
universal testing machine with a loading capacity of 1000 kN
was used to load specimens in compression. Pin-ended support
conditions were used for all specimens where the effective
length factor could be taken as 1.0. As suggested in the SSRC
guide [15], ball bearing supports were used to realize pinned
ends for each specimen. Each support consisted of two 13 mm
parallel plates separated by a stainless steel ball seated in upper
and lower sockets suitably milled into the plates as shown for
a base support in Fig. 3. Sockets in the bearing plates were
polished prior to each test to reduce friction with the steel balls.
The geometric centers of both top and bottom loading plates
were marked out and aligned with the steel ball and the center of
the loading head. For concentrically loaded angles, the centroid

Fig. 2. Test set-up.

of the angle cross-section was aligned with the center of the


plate assembly. For eccentrically loaded angles, the location of
the angle was determined and marked on the loading plate with
respect to the center of the plate. Fig. 4 shows a specimen that
had been aligned and tack welded to the supporting plate before
testing.

508

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

Fig. 5. LVDT locations for deflection measurements.


Fig. 3. Ball bearing support.

to ensure that the specimen and instrumentation were properly


seated in test position. During each test, loading was applied at
a constant rate of 6 kN per minute up to the failure while loads
and corresponding deflections were recorded at intervals of 1 s
using a computer-controlled data acquisition system. The test
was discontinued when the applied load dropped to 60% of the
ultimate load. Ultimate load was deemed to have been reached
when the specimen lost the ability to sustain the increasing load
while displaying a significant lateral deflection.
5. Test results

Fig. 4. Angle specimen welded on the bearing plate.

Both vertical deformation and lateral displacements were


measured during each test. Vertical deformation of a specimen
was measured with a built-in Linear Variable Differential
Transducer (LVDT) which measured travel of the loading head.
Lateral displacements were measured using LVDTs at midheight of each specimen and mid-width of each leg. The
bases of LVDTs were mounted on the two columns of the
Instron machine which travel vertically with the loading table.
Fig. 5 shows the lateral displacement measurement locations.
Displacement measurement locations of the LVDTs on the
angle legs changed as the twisting deflections began and
increased. Therefore additional measurements beyond the point
where the twisting began are not strictly valid and have been
used only to compare the advanced deflection responses of
specimens.
4. Test procedure
Prior to each test, the steel ball bearings were greased with
high pressure grease and care was taken to ensure that specimen
loading plate and steel ball were precisely aligned. The two
LVDTs were positioned to measure lateral deflections in
orthotropic directions at mid-height of the specimen. A nominal
compressive load of approximately 10% of the predicted
ultimate load was first applied and released. This procedure
was repeated twice more to remove slack from the system and

Table 1 lists the maximum load reached, P, and the observed


failure mode for each specimen. FB indicates flexural buckling
failure and TFB indicates a torsionalflexural buckling failure.
Nominally identical groups of specimens were specimens
11 and 12, specimens 14 and 15, and specimens 1820.
Specimens of each group were tested under nominally identical
loading conditions to verify repeatability of test procedures.
The variation of ultimate capacities in each group was within
10% and is considered to be within acceptable test limits. The
average capacity of the three slender specimens 1820 (E1500)
which were tested under concentric compression was 43.1 kN
as compared to a value of 45.9 kN based on equations suggested
in the SSRC guide [15]. This close agreement of test and
theoretical buckling loads is of further confirmation that the test
set-up and procedure were reliable and capable of providing
consistent test results within the acceptable limits of variation.
5.1. Effect of eccentricity
Normalized load P/Py vs. e/xo curves for specimens E900
and E1500 under compression causing both major-axis and
minor-axis bending are shown in Fig. 6. The ultimate load is
normalized using Py , the yield strength of the cross-section
(Py = AFy ) and the eccentricity is normalized using xo , the
distance between the shear center and the centroid of the crosssection as shown in Fig. 1. For L51 51 6.4, the value
of xo = 16.8 mm. Fig. 6 shows that specimens subjected
to minor-axis bending failed at a notably lower load than
those that failed under major-axis bending over the investigated
range of eccentricities for both slenderness ratios. While the
load capacity decreased with an increase in eccentricity, the

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

(a) E900 specimens.

509

(b) E1500 specimens.


Fig. 6. Normalized load P/Py vs. e/xo curves for E900 and E1500 specimens.

(a) Eccentric compression causing major-axis bending.

(b) Eccentric compression causing minor-axis bending.

Fig. 7. Normalized load P/Py vs. e/xo curves for specimens with two loading conditions.

decreasing trends are different depending on whether minoraxis or major-axis eccentricity prevails. As seen from Table 1
and Fig. 6a, the capacity reduction for E900 specimens as e/xo
varied from 0.0 to 0.6 was 10% for eccentric compression
causing major-axis bending. As e/xo varied from 0.0 to 0.5 a
reduction of 44% occurred for the case of minor-axis bending.
For E1500 specimens as e/xo varied from 0.0 to 0.5, the load
reduction was negligible for specimens subjected to majoraxis bending and 18% for specimens subjected to minoraxis bending. In fact, the variation of the ultimate load for
E1500 specimens subjected to major-axis bending remained
insignificant up to e/x0 = 1.75. It seems that for minor-axis
bending, the ultimate load is sensitive to the eccentricity within
the range of small eccentricities. As eccentricity increases, the
rate of decrease in ultimate load diminishes as indicated by the
flattening of the curves at large values of eccentricity.

5.2. Effect of slenderness ratio


The effect of slenderness ratio is illustrated in Fig. 7
where the normalized load, P/Py , is plotted against e/xo
for different slenderness ratios. As shown in Fig. 7a for
specimens under eccentric compression causing major-axis
bending, as the slenderness ratio increases, the ultimate
capacity decreases for all investigated eccentricities. The degree
of this decrease is associated with the value of eccentricity.
When the eccentricity is large, the decrease in capacity with
increasing slenderness becomes insignificant. On the other
hand, for a large slenderness ratio as in the case of E1500
specimens, the decrease in capacity with increasing eccentricity
becomes essentially negligible. It also shows that there seems
to exist a critical eccentricity for each slenderness ratio and
below this eccentricity, the ultimate load associated with majoraxis bending is not significantly affected by the variation of

510

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

(a) Eccentric compression causing major-axis bending.

(b) Eccentric compression causing minor-axis bending.

Fig. 8. Load vs. vertical deformation curves for E900 specimens.

eccentricity. Beyond this eccentricity, the reduction of the


capacity with an increase in eccentricity is more pronounced.
As shown in Fig. 7a, for specimens E900, E1200 and E1500,
the critical eccentricities are approximately 0.75 xo , 1.0 xo and
1.75 xo . For example, the reduction of the ultimate capacity of
specimens E900 as the eccentricity varied from 0 to 10 mm (0 to
0.75 xo ) was 10% in contrast to 19.7% as the eccentricity varied
from 10 to 16.6 mm (0.75 xo to 1.0 xo ). Similar observations
can be made for specimens E1200 and E1500 at eccentricity
of 1.0 xo and 1.75 xo . Specimen response due to minoraxis bending is summarized in Fig. 7b for specimens E900
and E1500 series. The trend towards decreasing capacity with
increasing eccentricity is evident here also. The rate of decrease
is much higher for the less slender E900 specimens than it
is for the more slender E1500 specimens. However, for both
slenderness ratios, no critical eccentricities similar to those in
the major-axis bending case can be identified.
5.3. Load vs. deflection curves
The load vs. vertical deformation behaviour for E900
specimens subjected to eccentric compression is shown in
Fig. 8a for major-axis bending and in Fig. 8b for minor-axis
bending. In Fig. 8a, as the eccentricity increases, the curves
show decreasing initial stiffnesses which is characteristic of
members subjected to eccentric compression. Also consistent
with this type of loading, as the eccentricity increases, the
nonlinear nature of the initial portion of the curves becomes
more pronounced as the corresponding ultimate load decreases.
Similar trends are seen in Fig. 8b for minor-axis bending.
The initial rate of decrease in ultimate capacity is quite
precipitous as it is influenced primarily by axial stresses at low
eccentricities. This precipitous decrease in ultimate capacity
gradually tapers off as eccentricity increases and bending
stresses begin to play a major role in the mode of failure for
both major- and minor-axis bending. Similar observations were
made during the study of load vs. vertical deformation response
for specimens of the E1200 and E1500 series.

Fig. 9. Comparison of load vs. vertical deformation curves for specimens of


three slenderness ratios (ex = 16.8 mm, e y = 0.0).

For comparison, load vs. vertical deformation curves are


shown in Fig. 9 for one specimen from each of the E900, E1200
and E1500 series with an applied load eccentricity of 16.8 mm
and having slenderness ratios, KL/ry, of 95, 125 and 155,
respectively. As expected, measured deflection in the axial load
direction of specimens loaded eccentrically begins as the load
increases from zero and continues at an increasing rate as the
stiffness decreases to zero at ultimate. Since these specimens
were not concentrically loaded, their ultimate capacities cannot
be compared solely on the basis of slenderness as used in
the simple Euler formula. However, at this eccentricity of
16.8 mm, it may be noted that a slenderness increase from 95
to 125 and from 125 to 155 resulted in decreases of ultimate
capacity of 18% and 40%, respectively. As noted above, the
rate of decrease in ultimate load with increasing slenderness
diminishes at higher eccentricities.
5.4. Failure mode
Flexural buckling and torsionalflexural buckling were
the observed failure modes. As indicated in Table 1, a
torsionalflexural buckling failure mode was observed only for

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

(a) Side view.

(b) Front view.

511

(c) Comparison of load vs. deflection curves for failure mode.


Fig. 10. Failure mode.

specimens subjected to major-axis bending. A flexural buckling


failure mode occurred in the cases where the specimens were
loaded concentrically, loaded eccentrically causing minor-axis
bending, or loaded at small eccentricities causing major-axis
bending.
In the majority of cases yielding bands began to form in
a region near mid-height of the E900 and E1200 specimens
prior to reaching ultimate load. The yielded region was less
apparent for the E1200 specimens and was not observed for
the E1500 specimens until well beyond ultimate. Post-failure
specimens, one from E1200 Series and one from E1500 Series,
are compared in Fig. 10a and b. Specimen E1200-4 failed by
torsionalflexural buckling and specimen E1500-2 failed by
flexural buckling. Fig. 10c compares the load vs. mid-height
lateral deflection curves of specimens E1200-4 and E15002. While the load vs. mid-height deflections for the two legs
of specimen E1500-2 are approximately symmetrical, those of
specimen E1200-4 clearly are not. These observations confirm
that specimen E1200-4 failed in a torsionalflexural buckling
mode whereas E1500-2 failed by flexural buckling. Similar
observations were used to confirm the failure modes of all other
specimens.
6. Comparison with available design methods
Test results were compared with design values suggested
in AISC Specification 2000, AISC Specification 2005 and
the modified equations proposed by Adluri and Madugula [5]
for single angles. AISC Specification 2005 for structural
steel buildings recommends that the in-plane and out-of-plane
stability be checked separately for a doubly symmetric member
in flexure and compression having applied moments primarily
in one plane. The in-plane stability equation is the same as Eq.

(1) whereas the out-of-plane buckling equation is expressed as


follows:


Mr 2
Pr
+
1.0
(3)
Pco
Mcx
where Pr and Mr are as defined in Eq. (1), Pco is the
available out-of-plane compressive strength, and Mcx is the
available flexuraltorsional strength for strong axis bending.
AISC Specification 2005 suggests that this equation provides
a better estimate of buckling loads where the buckling mode
is flexuraltorsional for doubly symmetric sections under
strong axis flexure. In the experimental study presented herein,
flexuraltorsional buckling was the predominant failure mode
observed for angle specimens under major-axis bending.
Table 2 presents a comparison of tested ultimate capacity,
P, and values obtained from AISC Specification 2000, PA2000 ,
from Adluri and Madugula procedure, PAM , and from AISC
Specification 2005, PA2005 . In the calculation of the suggested
design values, experimentally determined yield stress and effective specimen length with nominal cross-section dimensions
were used. Resistance factors were not included to facilitate
a comparison of raw experimental data and design-based values. In applying equations from all three procedures, interaction equations were checked at the three points labeled 1, 2
and 3 as shown in Fig. 1 and the resulting least load calculated is presented in Table 2 as the ultimate load. Overall, the
ratios of experimental ultimate loads to calculated design values ranged from 0.95 to 1.53. The average ratios of P/PA2000 ,
P/PAM and P/PA2005 are 1.22, 1.15 and 1.09 with C.O.Vs of
12.4%, 10.1%, and 8.0%, respectively. For specimens subjected
to eccentric compression causing minor-axis bending only,
AISC Specifications 2005 and 2000 equations yielded identical

512

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

Table 2
Comparison of test results with design estimates
No.

Specimen ID#

ex (mm)

e y (mm)

P (kN)

PA2000 (kN)

P
PA2000

PAM (kN)

P
PAM

PA2005 (kN)

P
PA2005

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28

E900-1
E900-2
E900-3
E900-4
E900-5
E900-6
E900-7
E900-8
E900-9
E900-10
E1200-1a
E1200-1b
E1200-2
E1200-3a
E1200-3b
E1200-4
E1200-5
E1500-1a
E1500-1b
E1500-1c
E1500-2
E1500-3
E1500-4
E1500-5
E1500-6
E1500-7
E1500-7
E1500-7

0.0
4.2
10.0
16.8
29.4
50.4
0.0
0.0
0.0
0.0
4.2
4.2
10.0
16.8
16.8
29.4
50.4
0.0
0.0
0.0
8.4
16.8
29.4
50.4
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
8.4
16.8
29.4
50.4
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
8.4
16.8
29.4
50.4

117.0
110.9
105.3
84.5
59.0
46.1
65.6
46.4
34.4
23.3
68.1
69.7
66.7
66.4
71.3
57.0
42.5
45.3
41.1
42.8
42.9
41.3
40.1
37.0
35.2
27.8
23.4
18.0

106.5
91.0
77.2
65.7
51.9
38.7
55.6
41.5
31.6
21.8
60.0
59.7
53.2
47.3
46.7
38.6
31.0
43.1
43.2
43.1
37.6
33.6
29.0
23.8
28.4
23.4
19.5
15.1
Ave.
C.O.V (%)

1.10
1.22
1.36
1.29
1.14
1.19
1.18
1.12
1.09
1.07
1.13
1.17
1.25
1.40
1.53
1.48
1.37
1.05
0.95
0.99
1.14
1.23
1.38
1.55
1.24
1.19
1.20
1.19
1.22
12.4

106.5
94.2
82.7
72.3
59.1
42.6
61.4
42.4
32.0
22.0
61.6
61.2
56.0
50.8
50.3
42.9
35.5
43.1
43.2
43.1
38.9
35.5
31.5
26.7
30.2
25.6
21.8
15.1

1.10
1.18
1.27
1.17
1.00
1.08
1.07
1.09
1.08
1.06
1.11
1.14
1.19
1.31
1.42
1.33
1.20
1.05
0.95
0.99
1.10
1.16
1.27
1.38
1.17
1.09
1.07
1.19
1.15
10.1

106.5
104.2
95.4
82.8
63.6
44.6
55.6
41.5
31.6
21.8
66.2
65.6
63.4
62.8
58.5
48.9
36.9
43.1
43.2
43.1
42.5
40.4
36.2
29.5
28.4
23.4
19.5
15.1

1.10
1.06
1.10
1.02
0.93
1.03
1.18
1.12
1.09
1.07
1.03
1.06
1.05
1.06
1.22
1.17
1.15
1.05
0.95
0.99
1.01
1.02
1.11
1.25
1.24
1.19
1.20
1.19
1.09
8.0

predicted capacities. When only specimens subjected to eccentric compression causing major-axis bending are considered,
the average ratios of P/PA2000 , P/PAM, and P/PA2005 are
1.31, 1.21 and 1.08 with C.O.Vs of 10.4%, 9.4%, and 7.8%, respectively. This suggests that the procedure suggested by Adluri
and Madugula [5] improved the average test-to-prediction by
7% overall, but the C.O.V. remained practically the same. Although intended for doubly symmetric members, AISC Specification 2005 equations provide improved test-to-prediction ratios and improved C.O.Vs for singly symmetric angles, especially for those subjected to major-axis bending. This suggests
that Eq. (3) can also be applied to singly symmetric sections
where the failure is flexuraltorsional buckling.
In Figs. 1113 for specimens subjected to eccentric
compression causing major-axis bending test results are further
compared with normalized load vs. normalized moment curves
obtained from AISC Specification 2000. Similar comparisons
are presented in Figs. 14 and 15 for specimens subjected to
eccentric compression causing minor-axis bending. M y is the
yield moment about the axis of bending and is calculated using
equations taken from AISC Specification 2000. Referring to the
cited figures and Table 2, it can be seen that the AISC 2000
provides conservative capacity estimates with the conservatism
being more significant in the case of specimens subjected
to major-axis bending than for those subjected to minoraxis bending. For specimens subjected to major-axis bending,

Fig. 11. Comparison of results and AISC Specification 2000 loadmoment


interaction curve for major-axis bending (Slenderness ratio = 95).

the underestimation of values using AISC Specification 2000


equations is increasingly pronounced as the slenderness ratio
increases. In the case of slenderness ratios, 125 and 155,
the predicted capacities are as low as about 50% of tested
values for some specimens. This marked underestimation for
specimens subjected to major-axis bending is attributed to
the fact that AISC Specification 2000 does not consider the
flexuraltorsional buckling behaviour.
7. Conclusions
The results of an experimental study involving the testing
of twenty-eight L51 51 6.4 single angles with slenderness

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

513

Fig. 12. Comparison of results and AISC Specification 2000 loadmoment


interaction curve for major-axis bending (Slenderness ratio = 125).
Fig. 15. Comparison of results and AISC Specification 2000 loadmoment
interaction curve for minor axis bending (Slenderness ratio = 155).

Fig. 13. Comparison of results and AISC Specification 2000 loadmoment


interaction curve for major-axis bending (Slenderness ratio = 155).

eccentric compression causing major-axis bending, a critical


eccentricity seems to exist within which, the reduction in
the ultimate load is insignificant. This critical eccentricity
assumes an increasingly larger value as the slenderness ratio
increases. However, this critical eccentricity is not observed
for specimens subject to eccentric compression causing minoraxis bending. When compared with experimental values, both
the AISC Specification 2000 and the procedure proposed by
Adluri and Madugula underestimate the axial load carrying
capacity of specimens under eccentric compression causing
major-axis bending. On the other hand, these methods provide
satisfactory results in the case of concentric loading and in the
case of eccentric loading causing minor-axis bending. AISC
Specification 2005 equations intended for doubly symmetric
sections provide improved capacity estimates of eccentrically
loaded single angles failing by flexuraltorsional buckling.
Acknowledgement
The authors wish to thank the Steel Structures Education
Foundation of Canada (SSEF) for financial support.
References

Fig. 14. Comparison of results and AISC Specification 2000 loadmoment


interaction curve for minor axis bending (Slenderness ratio = 95).

ratios of 95, 125, and 155 and subjected to loads of various


eccentricities, have been presented. Experimental results have
been compared with values obtained from AISC Specification
2000, AISC Specification 2005 and a procedure proposed
by Adluri and Madugula. The conclusions derived from this
research are listed as follows:
For nominally identical specimens with pinned end
boundary conditions, an increase in eccentricity results in
a reduction in ultimate load. For specimens subjected to

[1] ASCE 10-97. Design of latticed steel transmission structures. 1801


Alexander Bell Dr. (Reston, VA): American Society of Civil Engineers;
2000.
[2] CSA S37-01. Antennas, towers, and antenna-supporting structures.
Rexdale (Ontario): Canadian Standards Association; 2001.
[3] Load and resistance factor design specification for single-angle members.
Chicago (IL): American Institute of Steel Construction; 2000.
[4] Specification for structural steel buildings. Chicago (IL): American
Institute of Steel Construction; 2005.
[5] CAN/CSA S16-01. Handbook of steel construction. Mississauga
(Ontario, Canada): Canadian Institute of Steel Construction; 2004.
[6] Adluri SMR, Madugula MKS. Eccentrically loaded single angle struts.
Engineering Journal, AISC 1992;29(2):5966.
[7] Manual of steel construction-load and resistance factor design. 1st ed.
Chicago (IL): American Institute of Steel Construction; 1986.
[8] Wakabayashi M, Nonaka T. On the buckling strength of angles in
transmission towers. Bulletin of the Disaster Prevention Research
Institute, Kyoto University 1965;15(91). Part 2:1-18.

514

Y. Liu, L. Hui / Journal of Constructional Steel Research 64 (2008) 505514

[9] Ishida A. Experimental study on column carrying capacity of SHY steel


angles, vol. 265. Yawata technical report. Tokyo (Japan): Yawata Iron and
Steel Co. Ltd.; 1968. p. 856482; 87613.
[10] Mueller WH, Erzurumlu H. Behaviour and strength of angles in
compression: An experimental investigation. Research report of CivilStructural Engineering. Oregon (USA): Division of Engineering and
Applied Science, Portland State University; 1983.
[11] Sakla SSS. Performance of the AISC LRFD specification in predicting the
capacity of eccentrically loaded single-angle struts. Engineering Journal,
AISC 2005;42(40):23946.
[12] Earls CJ. Design of single angles bent about the major principal axis.
Engineering Journal, AISC 2003;40(3):15966.
[13] Madugula MKS, Kojima T, Kajita Y, Ohama M. Minor-axis bending
strength of angle beams. In: Proceedings of the international conference
on structural stability and design. 1995. p. 738.

[14] Earls CJ, Galambos TV. Practical compactness and bracing provisions for
the design of single angle beams. Engineering Journal, AISC 1998;35(1):
1925.
[15] SSRC guide. Guide to stability design criteria for metal structures.
Structural Stability Research Council. 5th ed. New York: John Wiley &
Sons, Inc.; 1998.
[16] Hui LB. Beamcolumn behaviour and strength of steel single equal-leg
angles. MASc thesis. Halifax (NS, Canada): Department of Civil and
Resource Engineering, Dalhousie University; 2007.
[17] Adluri SMR, Madugula MKS. Flexural buckling of steel angles:
Experimental investigation. Journal of Structural Engineering ASCE
1996;122(3):30917.
[18] ASTM A370-03a. Standard Methods and definitions for mechanical
testing of steel products, Annual Book of ASTM Standards, vol. 01.04.
2004.

Das könnte Ihnen auch gefallen